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.
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   

2.
We have recently reported that CoASH is the major low-molecular weight thiol in Bacillus anthracis [Nicely, N. I. , Parsonage, D., Paige, C., Newton, G. L., Fahey, R. C., Leonardi, R., Jackowski, S., Mallett, T. C., and Claiborne, A. (2007) Biochemistry 46, 3234-3245], and we have now characterized the kinetic and redox properties of the B. anthracis coenzyme A-disulfide reductase (CoADR, BACoADR) and determined the crystal structure at 2.30 A resolution. While the Staphylococcus aureus and Borrelia burgdorferi CoADRs exhibit strong preferences for NADPH and NADH, respectively, B. anthracis CoADR can use either pyridine nucleotide equally well. Sequence elements within the respective NAD(P)H-binding motifs correctly reflect the preferences for S. aureus and Bo. burgdorferi CoADRs, but leave questions as to how BACoADR can interact with both pyridine nucleotides. The structures of the NADH and NADPH complexes at ca. 2.3 A resolution reveal that a loop consisting of residues Glu180-Thr187 becomes ordered and changes conformation on NAD(P)H binding. NADH and NADPH interact with nearly identical conformations of this loop; the latter interaction, however, involves a novel binding mode in which the 2'-phosphate of NADPH points out toward solvent. In addition, the NAD(P)H-reduced BACoADR structures provide the first view of the reduced form (Cys42-SH/CoASH) of the Cys42-SSCoA redox center. The Cys42-SH side chain adopts a new conformation in which the conserved Tyr367'-OH and Tyr425'-OH interact with the nascent thiol(ate) on the flavin si-face. Kinetic data with Y367F, Y425F, and Y367,425F BACoADR mutants indicate that Tyr425' is the primary proton donor in catalysis, with Tyr367' functioning as a cryptic alternate donor in the absence of Tyr425'.  相似文献   

3.
An NAD(P)H-dependent H2O2 forming activity has been evidenced in thyroid tissue from patients with Grave's disease. Its biochemical properties were compared to those of the NADPH oxidase previously described in pig thyroid gland. Both were Ca2+-dependent and activated by inorganic phosphate anions in the same range of concentrations. Both are flavoproteins using FAD as cofactor, but the human enzyme was also able to utilize FMN. The apparent Km for NADPH of the human enzyme (100 microM) was 5-10 times higher than that of porcine enzyme. Vm was 3 to 10 times higher in pig (150 nmol x h(-1) x mg(-1)) than in man (14 to 45). Total content in human tissue was 7 to 9% of that in porcine tissue. An unidentified inhibitor has been detected in the 3000 g particulate fraction from most patients, which could account for this apparently low enzyme content. An NADH-dependent H2O2 production has also been observed in porcine and human thyroid tissues. This activity was only partly Ca2+-dependent (man, 50-70%; pig, 80-90%) and presented similar apparent Km values for NADH (man, 100 microM; pig, 200 microM). In pig thyrocytes, the expression of the Ca2+-dependent part of the NADH-oxidase activity was induced by TSH and down-regulated by TGFbeta, as was the NADPH oxidase activity. Furthermore, NADPH and NADH-dependent activities were not additive. We conclude that a single, inducible, NAD(P)H-oxidase can use NADPH or NADH as substrate to catalyse H2O2 formation, and that human and porcine NAD(P)H-oxidases are highly similar. Differences observed could be attributed to minor differences in enzyme structure and/or in membrane microenvironment. The NADH-dependent Ca2+-independent activity observed in human and porcine thyroid fractions could be attributed to a distinct and constitutive enzyme.  相似文献   

4.
Both the external oxidation of NADH and NADPH in intact potato (Solanum tuberosum L. cv. Bintje) tuber mitochondria and the rotenone-insensitive internal oxidation of NADPH by inside-out submitochondrial particles were dependent on Ca2+. The stimulation was not due to increased permeability of the inner mitochondrial membrane. Neither the membrane potential nor the latencies of NAD(+)-dependent and NADP(+)-dependent malate dehydrogenases were affected by the addition of Ca2+. The pH dependence and kinetics of Ca(2+)-dependent NADPH oxidation by inside-out submitochondrial particles were studied using three different electron acceptors: O2, duroquinone and ferricyanide. Ca2+ increased the activity with all acceptors with a maximum at neutral pH and an additional minor peak at pH 5.8 with O2 and duroquinone. Without Ca2+, the activity was maximal around pH 6. The Km for NADPH was decreased fourfold with ferricyanide and duroquinone, and twofold with O2 as acceptor, upon addition of Ca2+. The Vmax was not changed with ferricyanide as acceptor, but increased twofold with both duroquinone and O2. Half-maximal stimulation of the NADPH oxidation was found at 3 microM free Ca2+ with both O2 and duroquinone as acceptors. This is the first report of a membrane-bound enzyme inside the inner mitochondrial membrane which is directly dependent on micromolar concentrations of Ca2+. Mersalyl and dicumarol, two potent inhibitors of the external NADH dehydrogenase in plant mitochondria, were found to inhibit internal rotenone-insensitive NAD(P)H oxidation, at the same concentrations and in manners very similar to their effects on the external NAD(P)H oxidation.  相似文献   

5.
J L Gabriel  G W Plaut 《Biochemistry》1984,23(12):2773-2778
The activity of NAD-dependent isocitrate dehydrogenase from bovine heart was inhibited by NADH (apparent Ki about 4.3 microM) and NADPH (Ki about 9.8 microM) at subsaturating substrate concentrations at pH 7.4. The inhibition by NADH or NADPH was reversed competitively by magnesium isocitrate in the presence of ADP, but not without ADP. Reversal of inhibition by NADH or NADPH with respect to NAD+ was competitive or of the linear mixed type depending on whether ADP was absent or present. ADP3- (0.2 mM) increased the Ki(app) for NADPH from 9.8 to 27.1 microM; further addition of Ca2+ (0.2 mM) raised the Ki(app) to 127 microM. For the modification of NADPH inhibition by ADP, S0.5 for Ca2+ was approximately 48 microM. This compares to the Km for Ca2+ of 0.3-1 microM for the activation of the enzyme without NADPH [Denton, R. M., Richards, D. A., & Chin, J. G. (1978) Biochem. J. 176, 899-906; Aogaichi, T., Evans, J., Gabriel, J., & Plaut, G. W. E. (1980) Arch. Biochem. Biophys. 204, 350-360]. ADP did not affect the Ki for NADH. Magnesium citrate, which was about 100-fold more effective as a positive modifier of the enzyme with ADP than without ADP [Gabriel, J. L., & Plaut, G. W. E. (1983) Fed. Proc., Fed. Am. Soc. Exp. Biol. 42, 2082], reversed competitively the inhibition by NADPH in the presence of ADP, but not without ADP. Magnesium citrate did not reverse NADH inhibition.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Previous studies have shown that the reduced nicotinamide adenine dinucleotide phosphate (NADPH)- binding domain of rat liver microsomal steroid 5alpha-reductase isozyme-1 (r5alphaR-1) is in a highly conserved region of the polypeptide sequence (residues 160-190). In this study, we investigated, by site-directed mutagenesis, the role of hydroxylated and aromatic amino acids within the NADPH-binding domain. The r5alphaR-1 cDNA was cloned into a pCMV vector, and the double strand site-directed mutagenesis method was used to create mutants Y179F, Y179S, Y189F, Y189S, S164A, S164T, and Y187F, which were subsequently expressed in COS-1 cells. Kinetic studies of the expressed enzymes showed that the mutation Y179F resulted in an approximately 40-fold increase in the Km for NADPH versus wild-type, with only a 2-fold increase in the Km for testosterone. The mutants Y189F and S164A showed smaller increases (4 and 6-fold) in Kms for NADPH and no significant change in the Km for testosterone, whereas Y189S had kinetic properties similar to the wild-type r5alphaR-1. Mutants Y179S and S164T both resulted in inactive enzymes, whereas F187Y showed an approximately 5-fold decrease in Km for NADPH and a significant increase (approximately 18-fold) in the Km for testosterone. The results suggest that the -OH functionality of Y179 is involved in cofactor binding, but is not essential for the activity of the enzyme, whereas the -OH functionalities of Y189 and S164 play lesser roles in cofactor binding to r5alphaR-1 and may not be required for enzyme activity. On the other hand, the residue F187 may be important for the binding of both NADPH and testosterone.  相似文献   

7.
Activity staining with NADPH-nitroblue tetrazolium after native-PAGEof membrane proteins of Synechocystis PCC6803, solubilized with3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate (CHAPS),revealed four NAD(P)H dehydrogenase (NDH) activities; an NDHcomplex of the respiratory chain, a ferredoxin NADP+ reductase(FNR), a drgA product which oxidized both NADH and NADPH, andan uncharacterized NADH-specific enzyme. The NDH complex waspurified with anion exchange and gel filtration chromatographies.The purified complex had a molecular mass of 376 kDa and wascomposed of 9 subunits. Western analysis showed that the complexcontained the NDH-H subunit, but not NDH-A or B. The enzymereduced ferricyanide much faster than plastoquinone and usedNADPH as its prefered electron donor rather than NADH. The enzymaticactivity was inhibited by diphenyleneiodonium chloride and salicylhydroxamicacid, but not by rotenone, p-chloromercuribenzoate, N-ethylmaleimide,flavon, dicumarol, or antimycin A. These results suggest thatthe purified complex is a hydrophilic subcomplex which containsan NADPH binding site and flavin, and is dissociated from ahydrophobic subcomplex, which contains quinone binding site. 1Present address: Division of Applied Life Sciences, GraduateSchool of Agriculture, Kyoto University, Sakyo, Kyoto, 606-8502Japan 3Present address: Department of Biotechnology, Faculty of Engineering,Fukuyama University, 1 Gakuencho, Fukuyama, Hiroshima, 729-0292Japan  相似文献   

8.
1. After immunization of BALB/c mouse, four monoclonal antibodies against soluble NADH diaphorase from ejaculated boar spermatozoa were produced and characterized. The monoclonal antibodies were designated as follows Mab 1F2, Mab 4E2, Mab 5B8, Mab 5D8. 2. These monoclonal antibodies react with other enzyme forms-sedimentary NADH and NADPH and soluble NADPH and inhibit (although not completely) their activity. It is supposed that different forms of the enzyme share some common epitopes. 3. Treatment of ejaculated boar semen with 2O-methylcholanthrene causes an increase of the activity of the soluble diaphorase form only. 4. These results lead to the assumption that the sperm diaphorase is a dynamic enzyme system consisting of four immunologically similar isoenzymes although their functions are different.  相似文献   

9.
Quinone oxidoreductase activities dependent on pyridine nucleotides are associated with the plasma membrane (PM) in zucchini (Cucurbita pepo L.) hypocotyls. In the presence of NADPH, lipophilic ubiquinone homologs with up to three isoprenoid units were reduced by intact PM vesicles with a Km of 2 to 7 [mu]M. Affinities for both NADPH and NADH were similar (Km of 62 and 51 [mu]M, respectively). Two NAD(P)H:quinone oxidoreductase forms were identified. The first, labeled as peak I in gel-filtration experiments, behaves as an intrinsic membrane complex of about 300 kD, it slightly prefers NADH over NADPH, it is markedly sensitive to the inhibitor diphenylene iodonium, and it is active with lipophilic quinones. The second form (peak II) is an NADPH-preferring oxidoreductase of about 90 kD, weakly bound to the PM. Peak II is diphenylene iodonium-insensitive and resembles, in many properties, the soluble NAD(P)H:quinone oxidoreductase that is also present in the same tissue. Following purification of peak I, however, the latter gave rise to a quinone oxidoreductase of the soluble type (peak II), based on substrate and inhibitor specificities and chromatographic and electrophoretic evidence. It is proposed that a redox protein of the same class as the soluble NAD(P)H:quinone oxidoreductase (F. Sparla, G. Tedeschi, and P. Trost [1996] Plant Physiol. 112:249-258) is a component of the diphenylene iodonium-sensitive PM complex capable of reducing lipophilic quinones.  相似文献   

10.
Highly purified plasma membrane fractions were obtained from onion (Allium cepa L.) roots and used as a source for purification of redox proteins. Plasma membranes solubilized with Triton X-100 contained two distinct polypeptides showing NAD(P)H-dependent dehydrogenase activities. Dehydrogenase I was purified by gel filtration in Sephacryl S-300 HR, ion-exchange chromatography in DEAE-Sepharose CL-6B, and dye-ligand affinity chromatography in Blue-Sepharose CL-6B after biospecific elution with NADH. Dehydrogenase I consisted of a single polypeptide of about 27 kD and an isoelectric point of about 6. Dehydrogenase II was purified from the DEAE-unbound fraction by chromatography in Blue-Sepharose CL-6B and affinity elution with NADH. Dehydrogenase II consisted of a single polypeptide of about 31 kD and an isoelectric point of about 8. Purified dehydrogenase I oxidized both NADPH and NADH, although higher rates of electron transfer were obtained with NADPH. Maximal activity was achieved with NADPH as donor and juglone or coenzyme Q as acceptor. Dehydrogenase II was specific for NADH and exhibited maximal activity with ferricyanide. Optimal pH for both dehydrogenases was about 6. Dehydrogenase I was moderately inhibited by dicumarol, thenoyltrifluoroacetone, and the thiol reagent N-ethyl-maleimide. A strong inhibition of dehydrogenase II was obtained with dicumarol, thenoyltrifluoroacetone, and the thiol reagent p-hydroxymercuribenzoate.  相似文献   

11.
The soluble NAD(P)H:(quinone-acceptor) oxidoreductase [NAD(P)H-QR, EC 1.6.99.2] of Nicotiana tabacum L. leaves and roots has been purified. NAD(P)H-QR contains noncovalently bound flavin mononucleotide. Pairs of subunits of 21.4 kD are linked together by disulfide bridges, but the active enzyme is a homotetramer of 94 to 100 kD showing an isoelectric point of 5.1. NAD(P)H-QR is a B-stereospecific dehydrogenase. NADH and NADPH are electron donors of similar efficiency with Kcat:Km ratios (with duroquinone) of 6.2 x 107 and 8.0 x 107 m-1 s-1, respectively. Hydrophilic quinones are good electron acceptors, although ferricyanide and dichlorophenolindophenol are also reduced. The quinones are converted to hydroquinones by an obligatory two-electron transfer. No spectral evidence for a flavin semiquinone was detected following anaerobic photoreduction. Cibacron blue and 7-iodo-acridone-4-carboxylic acid are inhibitory. Tobacco NAD(P)H-QR resembles animal DT-diaphorase in some respects (identical reaction mechanism with a two-electron transfer to quinones, unusually high catalytic capability, and donor and acceptor substrate specificity), but it differs from DT-diaphorase in molecular structure, flavin cofactor, stereospecificity, and sensitivity to inhibitors. As in the case with DT-diaphorase in animals, the main NAD(P)H-QR function in plant cells may be the reduction of quinones to quinols, which prevents the production of semiquinones and oxygen radicals. The enzyme appears to belong to a widespread group of plant and fungal flavoproteins found in different cell compartments that are able to reduce quinones.  相似文献   

12.
Modulation of human CYP19A1 activity by mutant NADPH P450 oxidoreductase   总被引:1,自引:0,他引:1  
Mutations in NADPH P450 oxidoreductase (POR) cause a broad spectrum of human disease with abnormalities in steroidogenesis. We have studied the impact of P450 reductase mutations on the activity of CYP19A1. POR supported CYP19A1 activity with a calculated Km of 126 nm for androstenedione and a Vmax of 1.7 pmol/min. Mutations R457H and V492E located in the FAD domain of POR that disrupt electron transfer caused a complete loss of CYP19A1 activity. The A287P mutation of POR decreased the activities of CYP17A1 by 60-80% but had normal CYP19A1 activity. Molecular modeling and protein docking studies suggested that A287P is involved in the interaction of POR:CYP17A1 but not in the POR:CYP19A1 interaction. Mutations C569Y and V608F in the NADPH binding domain of POR had 49 and 28% of activity of CYP19A1 compared with normal reductase and were more sensitive to the amount of NADPH available for supporting CYP19A1 activity. Substitution of NADH for NADPH had a higher impact on C569Y and V608F mutants of POR. Similar effects were obtained at low/high (5.5/8.5) pH, but using octanol to limit the flux of electrons from POR to CYP19A1 inhibited activity supported by all variants. High molar ratios of KCl also reduced the CYP19A1 supporting activities of C569Y and V608F mutants of POR to a greater extent compared to normal POR and A287P mutant. Because POR supports many P450s involved in steroidogenesis, bone formation, and drug metabolism, variations in the effects of POR mutations on specific enzyme activities may explain the broad clinical spectrum of POR deficiency.  相似文献   

13.
Marohnic CC  Bewley MC  Barber MJ 《Biochemistry》2003,42(38):11170-11182
Microsomal cytochrome b(5) reductase (EC 1.6.2.2) catalyzes the reduction of ferricytochrome b(5) using NADH as the physiological electron donor. Site-directed mutagenesis has been used to engineer the soluble rat cytochrome b(5) reductase diaphorase domain to utilize NADPH as the preferred electron donor. Single and double mutations at residues D239 and F251 were made in a recombinant expression system that corresponded to D239E, S and T, F251R, and Y, D239S/F251R, D239S/F251Y, and D239T/F251R, respectively. Steady-state turnover measurements indicated that D239S/F251Y was bispecific while D239T, D239S/F251R, and D239T/F251R were each NADPH-specific. Wild-type (WT) cytochrome b(5) reductase showed a 3700-fold preference for NADH whereas the mutant with the highest NADPH efficiency, D239T, showed an 11-fold preference for NADPH, a 39200-fold increase. Wild-type cytochrome b(5) reductase only formed a stable charge-transfer complex with NADH while D239T formed complexes with both NADH and NADPH. The rates of hydride ion transfer, determined by stopped-flow kinetics, were k(NADH-WT) = 130 s(-1), k(NADPH-WT) = 5 s(-1), k(NADH-D239T) = 180 s(-1), and k(NADPH-D239T) = 73 s(-1). K(s) determinations by differential spectroscopy demonstrated that D239T could bind nonreducing pyridine nucleotides with a phosphate or a hydroxyl substituent at the 2' position, whereas wild-type cytochrome b(5) reductase would only bind 2' hydroxylated molecules. Oxidation-reduction potentials (E degrees ', n = 2) for the flavin cofactor were WT = -268 mV, D239T = -272 mV, WT+NAD(+) = -190 mV, D239T+NAD(+) = -206 mV, WT+NADP(+) = -253 mV, and D239T+NADP(+) = -215 mV, which demonstrated the thermodynamic contribution of NADP(+) binding to D239T. The crystal structures of D239T and D239T in complex with NAD(+) indicated that the loss of the negative electrostatic surface that precluded 2' phosphate binding in the wild-type enzyme was primarily responsible for the observed improvement in the use of NADPH by the D239T mutant.  相似文献   

14.
Ferredoxin-NADP(+) reductase catalyses NADP(+) reduction, being specific for NADP(+)/H. To understand coenzyme specificity determinants and coenzyme specificity reversion, mutations at the NADP(+)/H pyrophosphate binding and of the C-terminal regions have been simultaneously introduced in Anabaena FNR. The T155G/A160T/L263P/Y303S mutant was produced. The mutated enzyme presents similar k(cat) values for NADPH and NADH, around 2.5 times slower than that reported for WT FNR with NADPH. Its K(m) value for NADH decreased 20-fold with regard to WT FNR, whereas the K(m) for NADPH remains similar. The combined effect is a much higher catalytic efficiency for NAD(+)/H, with a minor decrease of that for NADP(+)/H. In the mutated enzyme, the specificity for NADPH versus NADH has been decreased from 67,500 times to only 12 times, being unable to discriminate between both coenzymes. Additionally, giving the role stated for the C-terminal Tyr in FNR, its role in the energetics of the FAD binding has been analysed.  相似文献   

15.
A freshwater Pseudomonas sp. was grown in continuous culture under steady-state conditions in L-lactate-, succinate-, glucose- or ammonium-limited media. Under carbon limitation, the NAD(H) (i.e. NAD + NADH) concentration of the organisms increased exponentially from approximately 2 to 7 mumol/g dry wt as the culture dilution rate (D) was decreased from 0.5 to 0.02 h-1. Organisms grown at a given D in any of the carbon-limited media possessed very similar levels of NAD(H). Therefore, under these conditions, cellular NAD(H) was only a function of the culture O and was independent of the nature of the culture carbon source. D had no influence on the NAD(H) content of cells grown under ammonium limitation. In contrast, cellular NADH concentration was not influenced by D in carbon- or ammonium-limited media. In L-lactate-limited medium, bacteria possessed 0.14 mumol NADH/g dry wt; very similar levels were found in organisms grown in the other media. The results are consistent with those of Wimpenny & Firth (1972) that bacteria rigidly maintain a constant NADH level rather than a constant constant NADH: NAD ratio. NADP(H) (i.e. NADP + NADPH) and NADPH levels were also not influenced by changes in the culture carbon source or in D; in L-lactate-limited medium these concentrations were 0.97 and 0.53 mumol/g cell dry wt, respectively. The NADPH:NADP(H) ratio was much higher than the NADH:NAD(H) ratio, averaging 55% in carbon-limited cells.  相似文献   

16.
A novel reductase displaying high specificity for glyoxylate and NADPH was purified 3343-fold from spinach leaves. The enzyme was found to be an oligomer of about 125 kDa, composed of four equal subunits of 33 kDa each. A Km for glyoxylate was about 14-fold lower with NADPH than with NADH (0.085 and 1.10 mM respectively), but the maximal activity, 210 mumol/min per mg of protein, was similar with either cofactor. Km values for NADPH and NADH were 3 and 150 microM respectively. Optimal rates with either NADPH or NADH were found in the pH range 6.5-7.4. The enzyme also showed some reactivity towards hydroxypyruvate with rates less than 2% of those observed for glyoxylate. Results of immunological studies, using antibodies prepared against either glyoxylate reductase or spinach peroxisomal hydroxypyruvate reductase, suggested substantial differences in molecular structure of the two proteins. The high rates of NADPH(NADH)-glyoxylate reductase in crude leaf extracts of spinach, wheat and soya bean (30-45 mumol/h per mg of chlorophyll) and its strong affinity for glyoxylate suggest that the enzyme may be an important side component of photorespiration in vivo. In leaves of nitrogen-fixing legumes, this reductase may also be involved in ureide breakdown, utilizing the glyoxylate produced during allantoate metabolism.  相似文献   

17.
R67 dihydrofolate reductase (R67 DHFR) catalyzes the transfer of a hydride ion from NADPH to dihydrofolate, generating tetrahydrofolate. The homotetrameric enzyme provides a unique environment for catalysis as both ligands bind within a single active site pore possessing 222 symmetry. Mutation of one active site residue results in concurrent mutation of three additional symmetry-related residues, causing large effects on binding of both ligands as well as catalysis. For example, mutation of symmetry-related tyrosine 69 residues to phenylalanine (Y69F), results in large increases in Km values for both ligands and a 2-fold rise in the kcat value for the reaction (Strader, M. B., Smiley, R. D., Stinnett, L. G., VerBerkmoes, N. C., and Howell, E. E. (2001) Biochemistry 40, 11344-11352). To understand the interactions between specific Tyr-69 residues and each ligand, asymmetric Y69F mutants were generated that contain one to four Y69F mutations. A general trend observed from isothermal titration calorimetry and steady-state kinetic studies of these asymmetric mutants is that increasing the number of Y69F mutations results in an increase in the Kd and Km values. In addition, a comparison of steady-state kinetic values suggests that two Tyr-69 residues in one half of the active site pore are necessary for NADPH to exhibit a wild-type Km value. A tyrosine 69 to leucine mutant was also generated to approach the type(s) of interaction(s) occurring between Tyr-69 residues and the ligands. These studies suggest that the hydroxyl group of Tyr-69 is important for interactions with NADPH, whereas both the hydroxyl group and hydrophobic ring atoms of the Tyr-69 residues are necessary for proper interactions with dihydrofolate.  相似文献   

18.
Essentially chlorophyll-free mitochondria were isolated from green leaves of spinach (Spinacia oleracea L. cv. Viking II). Uncoupled oxidation of exogenous NADPH (1 mM) to oxygen had an optimum at pH 6.0, and activity was relatively low at pH 7.0, even in the presence of 1 mM-CaCl2. There was a proportional increase in the apparent Km for NADPH with decreasing H+ concentrations, suggesting that NADPH protonated on the 2'-phosphate group was the true substrate. Exogenous NADH was oxidized by oxygen with an optimum at pH 6.9. Under low-cation conditions, EGTA or EDTA (both 1 mM) had no effect on the Vmax. of NADH oxidation, although the removal of bivalent cations from the membrane surface by the chelators could be observed by use of 9-aminoacridine fluorescence. In contrast, under high-cation conditions, chelators lowered the Vmax. by about 50%, probably due to a better approach of the negatively charged chelators to the negative membrane surface than under low-cation conditions. In a low-cation medium, the Vmax. of NADH oxidation was increased by about 50% by the addition of cations. This was caused by a lowering of the size of the negative surface potential through charge screening. In contrast with other cations, La3+ inhibited NADH oxidation, possibly through binding to lipids essential for NADH oxidation. The apparent Km for NADH varied 6-fold in response to changes in the size of the surface potential, suggesting that the approach of the negatively charged NADH to the active site is hampered by the negative surface potential. The results demonstrate that the spinach leaf cell can regulate the mitochondrial NAD(P)H oxidation through several mechanisms: the pH; the cation concentration in general; and the concentration of Ca2+ in particular. The results also emphasize the importance of electrostatic considerations when investigating the kinetic behaviour of membrane-bound enzymes.  相似文献   

19.
A yeast growing at 48°C was isolated from soil and the strain was identified as Cryptococcus lactativorus. The aldose reductase which the strain produced was purified 114-fold with an overall recovery of 36%. The stability of the enzyme was higher than that of other aldose reductases. The half life of the enzyme was 800 h and 14 h at 30°C and 50°C, respectively. The enzyme showed the best activity with d-xylose. l-Sorbose and d-fructose were also reduced by the enzyme. The enzyme was active with both NADPH and NADH as a conenzyme, and the activity with NADH was 1.25 times higher than that with NADPH. The Kmapp value for d-xylose was 8.6 mM and the Vmaxapp was 20.8 units/mg NADH was used as a coenzyme. The Kmapp values for NADPH and NADH were 6μM and 170 μM, respectively, when d-glucose was used as a substrate.  相似文献   

20.
The FAD-dependent NAD(P)H oxidase from Lactobacillus sanfrancisensis (L.san-Nox2) catalyzes the oxidation of 2 equivalents of either NADH or NADPH and reduces 1 equivalent of O(2) to yield 2 equivalents of water. During steady-state turnover only 0.5% of the reducing equivalents are detected in solution as hydrogen peroxide, suggesting that it is not released from the enzyme after the oxidation of the first equivalent of NAD(P)H and reaction with O(2). Here we report the crystal structure of L.san-Nox2 to 1.8 A resolution. The enzyme crystallizes as a dimer with each monomer consisting of a FAD binding domain (residues 1-120), a NAD(P)H binding domain (residues 150-250), and a dimerization domain (residues 325-451). The electron density for the redox-active Cys42 residue located adjacent to the si-face FAD is consistent with oxidation to the sulfenic acid (Cys-SOH) state. The side chain of Cys42 is also observed in two conformations; in one the sulfenic acid is hydrogen bonded to His10 and in the other it hydrogen bonds with the FAD O2' atom. Surprisingly, the NAD(P)H binding domains each contain an ADP ligand as established by electron density maps and MALDI-TOF analysis of the ligands released from heat-denatured enzyme. The ADP ligand copurifies with the enzyme, and its presence does not inhibit enzyme activity. Consequently, we hypothesize that either NADPH or NADH substrates bind via a long channel that extends from the enzyme exterior and terminates at the FAD re-face. A homology model of the NADH oxidase from Lactococcus lactis (L.lac-Nox2) was also generated using the crystal structure of L.san-Nox2, which reveals several important similarities and differences between the two enzymes. HPLC analysis of ligands released from denatured L.lac-Nox2 indicates that it does not bind ADP, which correlates with the specificity of the enzyme for oxidation of NADH.  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