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.
NAD(+)-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH), a member of the short-chain dehydrogenase family, catalyzes the first step in the catabolic pathway of the prostaglandins. This enzyme oxidizes the 15-hydroxyl group of prostaglandins to produce 15-keto metabolites which are usually biologically inactive. A relatively conserved threonine residue corresponding to threonine 11 of 15-PGDH is proposed to be involved in the interaction with NAD(+). Site-directed mutagenesis was used to examine the important role of this residue. Threonine 11 was changed to alanine (T11A), cysteine (T11C), serine (T11S) or tyrosine (T11Y) and the mutant proteins were expressed in E. coli. Western-blot analysis showed that the expression levels of mutant proteins were comparable to that of the wild-type enzyme. Mutants T11A, T11C and T11Y were found to be inactive. Mutant T11S still retained substantial activity and the K(m) value for prostaglandin E(2) (PGE(2)) was similar to the wild-type enzyme; however, the K(m) value for NAD(+) was increased over 23-fold. These results suggest that threonine 11 may be involved in the interaction with NAD(+) either directly or indirectly and contributes to the full catalytic activity of 15-PGDH.  相似文献   

2.
NAD+-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH) catalyzes the oxidation of the 15(S) hydroxyl group of prostaglandins to a 15-keto group resulting in a significant reduction of the biological activities of prostaglandins. Although the key residues involved in NAD+ binding and in catalytic activity have been partially identified, the sites of interaction of the enzyme with the prostaglandin substrates are yet to be determined. Homology analysis of the primary structures of 15-PGDH from human, mouse and rat indicates that the sequences are almost homologous except for two regions near the C-terminus. The involvement of the C-terminal region in catalytic activity was examined by studies on C-terminally truncated enzymes and on human/rat chimeric enzymes. When three to four amino acids were removed successively from the C-terminal end of human 15-PGDH, the truncated enzymes exhibited decreasing Vmax/Km ratios and increasing Km values for PGE2 as the chain was shortened. Similarly, when the C-terminal 14 amino acids of human 15-PGDH were replaced by the C-terminal 14 amino acids of rat 15-PGDH or vice versa, the Vmax/Km ratios and the Km values for prostaglandin E2 of the chimeric enzymes were in between those of the two wild-type enzymes. This indicates that the catalytic effectiveness of human 15-PGDH decreases as the C-terminal region is gradually removed or replaced by rat sequences. The C-terminal region appears to be more important for the interaction of the enzyme with the prostaglandin substrates than with the coenzyme.  相似文献   

3.
NAD+-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH), a member of the short chain dehydrogenase/reductase (SDR) family, is responsible for the biological inactivation of prostaglandins. Sequence alignment within SDR coupled with molecular modeling analysis has suggested that Gln-15, Asp-36, and Trp-37 of 15-PGDH may determine the coenzyme specificity of this enzyme. Site-directed mutagenesis was used to examine the important roles of these residues. Several single mutants (Q15K, Q15R, W37K, and W37R), double mutants (Q15K-W37K, Q15K-W37R, Q15R-W37K, and Q15R-W37R), and triple mutants (Q15K-D36A-W37R and Q15K-D36S-W37R) were prepared and expressed as glutathione S-transferase (GST) fusion proteins in Escherichia coli and purified by GSH-agarose affinity chromatography. Mutants Q15K, Q15R, W37K, W37R, Q15K-W37K, and Q15R-W37K were found to be inactive or almost inactive with NADP+ but still retained substantial activity with NAD+. Mutant Q15K-W37R and mutant Q15R-W37R showed comparable activity for NAD+ and NADP+ with an increase in activity nearly 3-fold over that of the wild type. However, approximately 30-fold higher in K(m) for NADP+ than that of the wild type enzyme for NAD+ was found for mutants Q15K-W37R and Q15R-W37R. Similarly, the K(m) values for PGE(2) of mutants were also shown to increase over that of the wild type. Further mutation of Asp-36 to either an alanine or a serine of the double mutant Q15K-W37R (i.e., triple mutants Q15K-D36A-W37R and Q15K-D36S-W37R) rendered the mutants exhibiting exclusive activity with NADP+ but not with NAD+. The triple mutants showed a decrease in K(m) for NADP+ but an increase in K(m) for PGE(2). Further mutation at Ala-14 to a serine of a triple mutant (Q15K-D36S-W37R) decreased the K(m) values for both NADP+ and PGE(2) to levels comparable to those of the wild type. These results indicate that the coenzyme specificity of 15-PGDH can be altered from NAD+ to NADP+ by changing a few critical residues near the N-terminal end.  相似文献   

4.
NAD(+)-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH), a member of the short-chain dehydrogenase/reductase (SDR) family, catalyzes the first step in the catabolic pathways of prostaglandins and lipoxins, and is believed to be the key enzyme responsible for the biological inactivation of these biologically potent eicosanoids. The enzyme utilizes NAD(+) specifically as a coenzyme. Potential amino acid residues involved in binding NAD(+) and facilitating enzyme catalysis have been partially identified. In this report, we propose that three more residues in 15-PGDH, Ile-17, Asn-91, and Val-186, are also involved in the interaction with NAD(+). Site-directed mutagenesis was used to examine their roles in binding NAD(+). Several mutants (I17A, I17V, I17L, I17E, I17K, N91A, N91D, N91K, V186A, V186I, V186D, and V186K) were prepared, expressed as glutathione S-transferase (GST) fusion enzymes in Escherichia coli, and purified by GSH-agarose affinity chromatography. Mutants I17E, I17K, N91L, N91K, and V186D were found to be inactive. Mutants N91A, N91D, V186A, and V186K exhibited comparable activities to the wild type enzyme. However, mutants I17A, I17V, I17L, and V186I had higher activity than the wild type. Especially, the activities of I17L and V186I were increased nearly 4- and 5-fold, respectively. The k(cat)/K(m) ratios of all active mutants for PGE(2) were similar to that of the wild type enzyme. However, the k(cat)/K(m) ratios of mutants I17A and N91A for NAD(+) were decreased 5- and 10-fold, respectively, whereas the k(cat)/K(m) ratios of mutants I17V, N91D, V186I, and V186K for NAD(+) were comparable to that of the wild type enzyme. The k(cat)/K(m) ratios of mutants I17L and V186A for NAD(+) were increased over nearly 2-fold. These results suggest that Ile-17, Asn-91, and Val-186 are involved in the interaction with NAD(+) and contribute to the full catalytic activity of 15-PGDH.  相似文献   

5.
NAD+-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH), a member of the short-chain dehydrogenase/reductase (SDR) family, catalyzes the first step in the catabolic pathways of prostaglandins and lipoxins. This enzyme oxidizes the C-15 hydroxyl group of prostaglandins and lipoxins to produce 15-keto metabolites which exhibit greatly reduced biological activities. A three-dimensional (3D) structure of 15-PGDH based on the crystal structures of the levodione reductase and tropinone reductase-II was generated and used for docking study with NAD+ coenzyme and PGE2 substrate. Three well-conserved residues among SDR family which correspond to Ser-138, Tyr-151, and Lys-155 of 15-PGDH have been shown to participate in the catalytic reaction. Based on the molecular interactions observed from 3D structure of 15-PGDH, we further propose that Gln-148 in 15-PGDH is important in properly positioning the 15-hydroxyl group of PGE2 by hydrogen bonding with the side-chain oxygen atom of Gln-148. This residue is found to be less conserved and replaceable by glutamyl, histidinyl, and asparaginyl residues in SDR family. Accordingly, site-directed mutagenesis of Gln-148 of 15-PGDH to alanine, glutamic acid, histidine, and asparagine (Q148A, Q148E, Q148H, and Q148N) was carried out. The activity of mutant Q148A was not detectable, whereas those of mutants Q148E, Q148H, and Q148N were comparable to or higher than the wild type. This indicates that the side-chain oxygen or nitrogen atom at position 148 of 15-PGDH plays an important role in anchoring C-15 hydroxyl group of PGE2 through hydrogen bonding for catalytic reaction.  相似文献   

6.
NAD(+)-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH) is a key enzyme involved in the catabolism of the prostaglandins. The cDNA for human placental 15-PGDH has been expressed in Escherichia coli as a catalytically active protein. The polymerase chain reaction was used to introduce restriction endonuclease sites at each end of the 15-PGDH coding sequence. The 15-PGDH DNA was then inserted into the bacterial expression plasmids pUC-18 and pUC-19 which contain the isopropyl-l-thio-beta-D-galactopyranoside (IPTG) inducible lacZ promoter. Extracts from E. coli containing these expression plasmids exhibited 15-PGDH activity which was inducible with (IPTG). Crude extracts from E. coli expressing 15-PGDH activity were found to contain proteins of the predicted sizes in stained SDS-polyacrylamide gels and in Western blots using human placental 15-PGDH antiserum. The specific activity in E. coli extracts was several hundred-fold higher than that seen in extracts from human placenta.  相似文献   

7.
15-Hydroxyprostaglandin dehydrogenase (15-PGDH) catalyzes NAD(+)-dependent oxidation of 15(S)-hydroxyl group of prostaglandins and has been considered a key enzyme involved in biological inactivation of prostaglandins. This enzyme is markedly induced by androgens in hormone-sensitive human prostate cancer cells (Tong M., Tai H. H. Biochem Biophys Res Commun 2000; 276: 77-81) and may be involved in tumorigenesis. Inhibition of this enzyme may be of value in anticancer therapy. Non-steroidal anti-inflammatory drugs (NSAIDs) which inhibit cyclooxygenases (COXs) have been shown to be chemopreventive in epidemiological and animal-model studies. However, chemoprevention by these drugs may not be directly related to their inhibition of COXs. Other targets may be also involved in their chemopreventive activity. We have examined a variety of NSAIDs including COX-2 selective inhibitors, peroxisome proliferator-activated receptor (PPAR) gamma agonists and phytophenolic compounds which have been shown to be chemopreventive for their effect on 15-PGDH. It was found that most of these compounds were potent inhibitors of 15-PGDH. Among these compounds, ciglitazone appeared to be the most powerful inhibitor (IC(50)=2.7 microM). Inhibition by ciglitazone was non-competitive with respect to NAD(+) and uncompetitive with respect to PGE(2).  相似文献   

8.
We previously reported the expression of a full-length cDNA complementary to a rat liver NAD(P)H:quinone oxidoreductase (EC 1.6.99.2) mRNA in Escherichia coli (Q. Ma, R. Wang, C. S. Yang, and A. Y. H. Lu, 1990, Arch. Biochem. Biophys. 283, 311-317). Since cysteine residues have been suggested to be important for the catalysis of flavoproteins and a lysine residue at position 76 in NAD(P)H:quinone oxidoreductase has been proposed to be involved in electron transfer of the enzyme, we investigated the roles of lysine 76 and cysteine 179 of this enzyme in catalysis by site-directed mutagenesis. Mutant cDNA clones replacing lysine 76 with valine (K76V) and cysteine 179 with alanine (C179A) were generated by a procedure based on the polymerase chain reaction. The mutant enzymes were expressed in E. coli. The cytosolic activities of the K76V and C179A mutants were 50 and 25% of that of the wild type (DTD), due to lower levels of the mutant proteins as shown by immunoblot analysis. The mutant proteins were purified to apparent homogeneity. The purified K76V and C179A mutant enzymes maintained full activities of 2,6-dichlorophenolindophenol (DCIP) reduction compared with that of the wild type. The mutant enzymes exhibited kinetic parameters for DCIP, NADH, and NADPH similar to those of DTD except that, with K76V, the Km for NADPH was doubled. Both mutant proteins contained two molecules of FAD per enzyme molecule. Dicumarol inhibited K76V and C179A mutant activities to greater than 90% at a concentration of 10(-7) M. Heat stability studies showed that C179A was much more sensitive to inactivation at 37 degrees C than both the wild-type and K76V enzymes. It is concluded from this study that lysine 76 and cysteine 179 are not essential in catalysis and in the binding of FAD, DCIP, and dicumarol. However, lysine residue 76 appears to play a role in NADPH binding and cysteine residue 179 is important in maintaining the stability of the enzyme.  相似文献   

9.
Recently, we amplified the Lactobacillus bulgaricus NAD(+)-dependent D-lactate dehydrogenase gene by the polymerase chain reaction, cloned and overexpressed it in Escherichia coli (Kochhar, S., Chuard, N., and Hottinger, H. (1992) Biochem. Biophys. Res. Commun. 185, 705-712). Polymerase chain reaction-amplified DNA fragments may contain base changes resulting in mutant gene products. A comparison of specific activities of D-lactate dehydrogenase in the crude extracts of 50 recombinant clones indicated that one of the clones had drastically reduced enzyme activity. Nucleotide sequence analysis of the insert DNA showed an exchange of A to G at position 795 resulting in substitution of Glu264 to Gly in the D-lactate dehydrogenase. The purified mutant D-lactate dehydrogenase showed a shift of 2 units in its optimum pH toward the acidic range. The dependence of kcat/Km on the pH of the mutant enzyme showed that the pKa of the free enzyme was around 4, at least 2 pH units lower than that of the wild-type enzyme. Both the wild-type and the mutant enzyme at their respective optimum pH values showed similar kcat and Km values. The data suggest that the highly conserved Glu264 is not critical for enzyme catalysis, but it must be situated within hydrogen bonding distance to amino acid residue(s) involved in substrate binding as well as in catalysis.  相似文献   

10.
The human mitochondrial NAD(P)+-dependent malic enzyme (m-NAD-ME) is a malic enzyme isoform with dual cofactor specificity and substrate binding cooperativity. Previous kinetic studies have suggested that Lys362 in the pigeon cytosolic NADP+-dependent malic enzyme has remarkable effects on the binding of NADP+ to the enzyme and on the catalytic power of the enzyme (Kuo, C. C., Tsai, L. C., Chin, T. Y., Chang, G.-G., and Chou, W. Y. (2000) Biochem. Biophys. Res. Commun. 270, 821-825). In this study, we investigate the important role of Gln362 in the transformation of cofactor specificity from NAD+ to NADP+ in human m-NAD-ME. Our kinetic data clearly indicate that the Q362K mutant shifted its cofactor preference from NAD+ to NADP+. The Km(NADP) and kcat(NADP) values for this mutant were reduced by 4-6-fold and increased by 5-10-fold, respectively, compared with those for the wild-type enzyme. Furthermore, up to a 2-fold reduction in Km(NADP)/Km(NAD) and elevation of kcat(NADP)/kcat(NAD) were observed for the Q362K enzyme. Mutation of Gln362 to Ala or Asn did not shift its cofactor preference. The Km(NADP)/Km(NAD) and kcat(NADP)/kcat(NAD) values for Q362A and Q362N were comparable with those for the wild-type enzyme. The DeltaG values for Q362A and Q362N with either NAD+ or NADP+ were positive, indicating that substitution of Gln with Ala or Asn at position 362 brings about unfavorable cofactor binding at the active site and thus significantly reduces the catalytic efficiency. Our data also indicate that the cooperative binding of malate became insignificant in human m-NAD-ME upon mutation of Gln362 to Lys because the sigmoidal phenomenon appearing in the wild-type enzyme was much less obvious that that in Q362K. Therefore, mutation of Gln362 to Lys in human m-NAD-ME alters its kinetic properties of cofactor preference, malate binding cooperativity, and allosteric regulation by fumarate. However, the other Gln362 mutants, Q362A and Q362N, have conserved malate binding cooperativity and NAD+ specificity. In this study, we provide clear evidence that the single mutation of Gln362 to Lys in human m-NAD-ME changes it to an NADP+-dependent enzyme, which is characteristic because it is non-allosteric, non-cooperative, and NADP+-specific.  相似文献   

11.
Yeast NAD(+)-specific isocitrate dehydrogenase (IDH) is an octamer containing two types of homologous subunits. Ligand-binding analyses were conducted to examine effects of residue changes in putative catalytic and regulatory isocitrate-binding sites respectively contained in IDH2 and IDH1 subunits. Replacement of homologous serine residues in either subunit site, S98A in IDH2 or S92A in IDH1, was found to reduce by half the total number of holoenzyme isocitrate-binding sites, confirming a correlation between detrimental effects on isocitrate binding and respective kinetic defects in catalysis and allosteric activation by AMP. Replacement of both serine residues eliminates isocitrate binding and measurable catalytic activity. The putative isocitrate-binding sites of IDH1 and IDH2 contain five identical and four nonidentical residues. Reciprocal replacement of the four nonidentical residues in either or both subunits (A108R, F136Y, T241D, and N245D in IDH1 and/or R114A, Y142F, D248T, and D252N in IDH2) was found to be permissive for isocitrate binding. This provides further evidence for two types of binding sites in IDH, although the authentic residues have been shown to be necessary for normal kinetic contributions. Finally, the mutant enzymes with residue replacements in the IDH1 site were found to be unable to bind AMP, suggesting that allosteric activation is dependent both upon binding of isocitrate at the IDH1 site and upon the changes in the enzyme normally elicited by this binding.  相似文献   

12.
Chloroplast glyceraldehyde-3-phosphate dehydrogenase (GAPDH) of higher plants uses both NADP(H) and NAD(H) as coenzyme and consists of one (GapA) or two types of subunits (GapA, GapB). AB-GAPDH is regulated in vivo through the action of thioredoxin and metabolites, showing higher kinetic preference for NADPH in the light than in darkness due to a specific effect on kcat(NADPH). Previous crystallographic studies on spinach chloroplast A4-GAPDH complexed with NADP or NAD showed that residues Thr33 and Ser188 are involved in NADP over NAD selectivity by interacting with the 2'-phosphate group of NADP. This suggested a possible involvement of these residues in the regulatory mechanism. Mutants of recombinant spinach GapA (A4-GAPDH) with Thr33 or Ser188 replaced by Ala (T33A, S188A and double mutant T33A/S188A) were produced, expressed in Escherichia coli, and compared to wild-type recombinant A4-GAPDH, in terms of crystal structures and kinetic properties. Affinity for NADPH was decreased significantly in all mutants, and kcat(NADPH) was lowered in mutants carrying the substitution of Ser188. NADH-dependent activity was unaffected. The decrease of kcat/Km of the NADPH-dependent reaction in Ser188 mutants resembles the behaviour of AB-GAPDH inhibited by oxidized thioredoxin, as confirmed by steady-state kinetic analysis of native enzyme. A significant expansion of size of the A4-tetramer was observed in the S188A mutant compared to wild-type A4. We conclude that in the absence of interactions between Ser188 and the 2'-phosphate group of NADP, the enzyme structure relaxes to a less compact conformation, which negatively affects the complex catalytic cycle of GADPH. A model based on this concept might be developed to explain the in vivo light-regulation of the GAPDH.  相似文献   

13.
The S1 subunit of pertussis toxin catalyses the hydrolysis of NAD+ (NAD+ glycohydrolysis) and the NAD(+)-dependent ADP-ribosylation of guanine-nucleotide-binding proteins. Recently, the S1 subunit of pertussis toxin was shown to be photolabelled by using radiolabelled NAD+ and u.v.; the primary labelled residue was Glu-129, thereby implicating this residue in the binding of NAD+. Studies from various laboratories have shown that the N-terminal portion of the S1 subunit, which shows sequence similarity to cholera toxin and Escherichia coli heat-labile toxin, is important to the maintenance of both glycohydrolase and transferase activity. In the present study the photolabelling technique was applied to the analysis of a series of recombinant-derived S1 molecules that possessed deletions or substitutions near the N-terminus of the S1 molecule. The results revealed a positive correlation between the extent of photolabelling with NAD+ and the magnitude of specific NAD+ glycohydrolase activity exhibited by the mutants. Enzyme kinetic analyses of the N-terminal mutants also identified a mutant with substantially reduced activity, a depressed photolabelling efficiency and a markedly increased Km for NAD+. The results support a direct role for the N-terminal region of the S1 subunit in the binding of NAD+, thereby providing a rationale for the effect of mutations in this region on enzymic activity.  相似文献   

14.
Homology modeling, molecular docking, and molecular dynamics simulation have been performed to determine human 15-hydroxyprostaglandin dehydrogenase (15-PGDH) binding with its NAD+ cofactor and prostaglandin E2 (PGE2) substrate. The computational studies have led to a three-dimensional (3D) model of the entire 15-PGDH-NAD+-PGE2 complex, demonstrating the detailed binding of PGE2 with 15-PGDH for the first time. This 3D model shows specific interactions of the protein with the cofactor and substrate in qualitative agreement with available experimental data. Our model demonstrates the PGE2-binding cavity of the protein for the first time. The model further leads to an interesting prediction that the catalytic activity of 15-PGDH should also significantly be affected by Gln148, in addition to the previously known three catalytic residues (Ser138, Tyr151, and Lys155). The reported 3D model of 15-PGDH-NAD+-PGE2 complex might be valuable for future rational design of novel inhibitors of 15-PGDH.  相似文献   

15.
Perdeuterated spin label (DSL) analogs of NAD+, with the spin label attached at either the C8 or N6 position of the adenine ring, have been employed in an EPR investigation of models for negative cooperativity binding to tetrameric glyceraldehyde-3-phosphate dehydrogenase and conformational changes of the DSL-NAD+-enzyme complex during the catalytic reaction. C8-DSL-NAD+ and N6-DSL-NAD+ showed 80 and 45% of the activity of the native NAD+, respectively. Therefore, these spin-labeled compounds are very efficacious for investigations of the motional dynamics and catalytic mechanism of this dehydrogenase. Perdeuterated spin labels enhanced spectral sensitivity and resolution thereby enabling the simultaneous detection of spin-labeled NAD+ in three conditions: (1) DSL-NAD+ freely tumbling in the presence of, but not bound to, glyceraldehyde-3-phosphate dehydrogenase, (2) DSL-NAD+ tightly bound to enzyme subunits remote (58 A) from other NAD+ binding sites, and (3) DSL-NAD+ bound to adjacent monomers and exhibiting electron dipolar interactions (8-9 A or 12-13 A, depending on the analog). Determinations of relative amounts of DSL-NAD+ in these three environments and measurements of the binding constants, K1-K4, permitted characterization of the mathematical model describing the negative cooperativity in the binding of four NAD+ to glyceraldehyde-3-phosphate dehydrogenase. For enzyme crystallized from rabbit muscle, EPR results were found to be consistent with the ligand-induced sequential model and inconsistent with the pre-existing asymmetry models. The electron dipolar interaction observed between spin labels bound to two adjacent glyceraldehyde-3-phosphate dehydrogenase monomers (8-9 or 12-13 A) related by the R-axis provided a sensitive probe of conformational changes of the enzyme-DSL-NAD+ complex. When glyceraldehyde-3-phosphate was covalently bound to the active site cysteine-149, an increase in electron dipolar interaction was observed. This increase was consistent with a closer approximation of spin labels produced by steric interactions between the phosphoglyceryl residue and DSL-NAD+. Coenzyme reduction (DSL-NADH) or inactivation of the dehydrogenase by carboxymethylation of the active site cysteine-149 did not produce changes in the dipolar interactions or spatial separation of the spin labels attached to the adenine moiety of the NAD+. However, coenzyme reduction or carboxymethylation did alter the stoichiometry of binding and caused the release of approximately one loosely bound DSL-NAD+ from the enzyme. These findings suggest that ionic charge interactions are important in coenzyme binding at the active site.  相似文献   

16.
Yeast NAD+-specific isocitrate dehydrogenase (IDH) is an octamer of four IDH1 and four IDH2 subunits, and the basic structural unit of the enzyme is an IDH1/IDH2 heterodimer. To investigate one aspect of the interaction between IDH1 and IDH2, residues in a hydrophobic region at the heterodimer interface (Val-216, Ser-220, and Val-224 in IDH1; Ile-221, Val-225, and Val-229 in IDH2) were replaced by alanine residues in each and in both subunits. Gel filtration and sedimentation velocity analyses demonstrated that the residue substitutions do not disrupt the octameric structure of IDH. However, these substitutions produce novel kinetic properties including, with respect to cofactor, positive allosteric regulation by AMP and cooperativity in the absence of AMP. These allosteric properties are also apparent in NAD+-binding experiments. Despite substantial measurable activity for the mutant enzyme containing residue substitutions in both subunits, expression of this enzyme produces growth phenotypes indicative of IDH dysfunction in vivo.  相似文献   

17.
Site-directed mutagenesis was utilized to identify binding sites for NAD(P)H and dicumarol in rat liver NAD(P)H:quinone oxidoreductase (NQOR, EC 1.6.99.2). The mutant cDNA clones were generated by a procedure based on the polymerase chain reaction and were expressed in Escherichia coli. The mutant enzymes were purified to apparent homogeneity as judged by SDS-polyacrylamide gel electrophoresis and were found to contain 2 FADs/enzyme molecule identical with that of the wild-type NQOR. Purified mutant enzymes Y128D, G150F, G150V, S151F, and Y155D showed dramatic decreases in activities in the reduction of dichlorophenolindophenol in comparison with the activities of the wild-type enzyme, whereas the activities of F124L, T127V, T127E, Y128V, Y128F, S151A, and Y155V were similar to those of NQOR. Enzyme kinetic analysis revealed that the Km values of T127E, Y128D, G150F, G150V, S151F, and Y155D were, respectively, 4-, 2-, 13-, 5-, 26-, and 19-fold higher than the Km of NQOR for NADPH, and were, respectively, 2-, 3-, 7-, 3-, 20-, and 11-fold higher than that of NQOR for NADH. The kcat values of Y128D, G150F, and G150V were also much lower than those of NQOR, but the kcat values of other mutants were similar to those of the wild-type enzyme. The Km values of the mutants for dichlorophenolindophenol were the same or slightly higher than that of NQOR. The apparent inhibition constants (Ki) for dicumarol on Y128V and F124L were elevated 12 and 8 times, respectively. Similar, but smaller, changes on Ki for 4-hydroxycoumarin were also observed. This study demonstrated that residues Gly150, Ser151, and Tyr155 in the glycine-rich region of NQOR are essential for NADPH and NADH binding and Tyr128 is important for dicumarol binding. Based on the results of the study, it is proposed that the glycine-rich region of the enzyme, along with other residues around the region, forms a beta sheet-turn-alpha helix structure important for the binding of the pyrophosphate group of NADPH and NADH.  相似文献   

18.
12-L-hydroxy-5,8,10-heptadecatrienoic acid (HHT) was found to be an excellent substrate for NAD+ dependent 15-hydroxyprostaglandin dehydrogenase from porcine kidney. Kcat/Km value of HHT was comparable to that of prostaglandin E although HHT is not a prostanoic acid derivative. Product of enzyme catalyzed oxidation of HHT was identified as 12-keto-5,8,10-heptadecatrienoic acid by gas chromatography-mass spectrometry. The fact that HHT is an excellent substrate for 15-hydroxyprostaglandin dehydrogenase suggest that HHT may have profound unrecognized biological actions and its inactivation may be via oxidation of the hydroxyl group.  相似文献   

19.
Erythroascorbic acid (eAsA) is a five-carbon analog of ascorbic acid, and it is synthesized from D-arabinose by D-arabinose dehydrogenase (ARA) and D-arabinono-gamma-lactone oxidase. We found an NAD+-specific ARA activity which is operative under submillimolar level of d-arabinose in the extracts of Saccharomyces cerevisiae. The hypothetical protein encoded by YMR041c showed a significant homology to a l-galactose dehydrogenase which plays in plant ascorbic acid biosynthesis, and we named it as Ara2p. Recombinant Ara2p showed NAD+-specific ARA activity with Km=0.78 mM to d-arabinose, which is 200-fold lower than that for the conventional NADP+-specific ARA, Ara1p. Gene disruptant of ARA2 lost entire NAD+-specific ARA activity and the conspicuous increase in intracellular eAsA by exogenous d-arabinose feeding, while the double knockout mutant of ARA1 and ARA2 still retained measurable amount of eAsA. It demonstrates that Ara2p, not Ara1p, mainly contributes to the production of eAsA from d-arabinose in S. cerevisiae.  相似文献   

20.
The cdc2 protein kinase family is regulated negatively by phosphorylation in the glycine ATP-binding loop at a conserved tyrosine residue, Y15, alone or in combination with T14 phosphorylation. In Schizosaccharomyces pombe and other systems, substitution of these residues with structurally similar but nonphosphorylatable amino acids has generated proteins (Y15F or T14AY15F) that behave as constitutively tyrosine-dephosphorylated proteins or threonine and tyrosine-dephosphorylated proteins. Here we report the characteristics of three additional mutants at Y15--Y15E, Y15S, and Y15T--in S. pombe cdc2p. All three mutant proteins are active in in vitro kinase assays, but are unable to functionally complement cdc2 loss-of-function mutations in vivo. Additionally, all three mutants are dominant negatives. A more detailed analysis of the Y15T mutant indicates that it can initiate chromosome condensation and F-actin contractile ring formation, but is unable to drive the reorganization of microtubules into a mitotic spindle.  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