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 78 毫秒
1.
Bandarian V  Reed GH 《Biochemistry》2002,41(27):8580-8588
The structure of the steady-state radical intermediate in the deamination of S-2-aminopropanol catalyzed by ethanolamine ammonia-lyase (EAL) from Salmonella typhimurium has been probed by electron paramagnetic resonance (EPR) spectroscopy using isotopically labeled forms of the substrate and of the adenosylcobalamin cofactor. Electron spin-spin coupling between the radical, centered on the carbon skeleton of the substrate, and the low-spin Co(2+) in cob(II)alamin (B(12r)) produces a dominant splitting of the EPR signals of both the radical and the Co(2+). Analysis of the exchange and dipole-dipole contributions to the spin-spin coupling indicates that the two paramagnetic centers are separated by approximately 11 A. Experiments with (13)C- and with (2)H-labeled forms of S-2-aminopropanol show that the radical is centered on C1 of the carbon skeleton of the substrate in agreement with an earlier report [Babior, B. M., Moss, T. H., Orme-Johnson, W. H., and Beinert, H., (1974) J. Biol. Chem. 249, 4537-4544]. Experiments with perdeutero-S-2-aminopropanol and [2-(15)N]-perdeutero-S-2-aminopropanol reveal a strong hyperfine splitting from the substrate nitrogen, which indicates that the radical is the initial substrate radical created by abstraction of a hydrogen atom from C1 of S-2-aminopropanol. The strong nitrogen hyperfine splitting further indicates that the amino substituent at C2 is approximately eclipsed with respect to the half-occupied p orbital at C1. Experiments with adenosylcobalamin enriched in (15)N in the dimethylbenzimidazole moiety show that the axial base of the cofactor remains attached to the Co(2+) in a functional steady-state reaction intermediate.  相似文献   

2.
Bandarian V  Reed GH 《Biochemistry》2000,39(39):12069-12075
Transient phases of the reaction catalyzed by ethanolamine ammonia-lyase (EAL) from Salmonella typhimurium have been investigated by stopped-flow visible spectrophotometry and deuterium kinetic isotope effects. The cleavage of adenosylcobalamin (coenzyme B(12)) to form cob(II)alamin (B(12r)) with ethanolamine as the substrate occurred within the dead time of the instrument whenever coenzyme B(12) was preincubated with enzyme prior to mixing with substrate. The rate was, however, slowed sufficiently to be measured with perdeutero ethanolamine as the substrate. Optical spectra indicate that, during the steady states of the reactions with ethanolamine and with S-2-aminopropanol as substrates, approximately 90% of the active sites contain B(12r). Reformation of the carbon-cobalt bond of the cofactor occurs following depletion of substrate in the reaction mixtures, and the rate constant for this process reflects k(cat) of the respective substrates. This late phase of the reaction also exhibits (2)H isotope effects similar to those measured for the overall reaction with (2)H-labeled substrates. With unlabeled substrates, the rate of cofactor reassembly is independent of the number of substrate molecules turned over in the steady-state phase. However, with (2)H-labeled substrates, kinetic isotope effects appear in the reassembly phase, and these isotope effects are maximal after only approximately 2 equiv of substrate/active site are processed. With 5'-deuterated coenzyme B(12) and deuterated substrate, the isotope effect on reassembly is independent of the number of substrate molecules that are turned over. These results indicate that the pool of exchangeable hydrogens in the enzyme-cofactor complex is two-a finding consistent with the hydrogens in the C5' methylene of coenzyme B(12).  相似文献   

3.
The non-functional form of xanthine oxidase known as the desulpho enzyme was compared with the functional enzyme in various ways, to obtain information on the structure of the molybdenum centre and the mechanism of the catalytic reaction. The desulpho enzyme, like the functional one, possesses a site for the binding of anions, presumably as ligands of molybdenum. Evidence is presented that in the Mo(V) e.p.r. signal from the desulpho-enzyme, as in that from the functional enzyme, a weakly coupled proton, in addition to a strongly coupled proton, interacts with the metal. Measurements were carried out by e.p.r. on the rate at which the proton strongly coupled to molybdenum exchanged, on diluting enzyme samples with 2H2O. For the desulpho enzyme the exchange rate constant was 0.40s-1, at pH 8.2 and 12 degrees C, and for the functional enzyme it was 85 s-1. It is shown that the great majority of reported differences between the enzyme forms are consistent with functional enzyme containing an (Enzyme)-Mo=S grouping, replaced in the desulpho form by (Enzyme)-Mo=O. Protonation of these groups, with pK values of about 8 and 10 respectively, would give (Enzyme)-Mo-SH and (Enzyme)-Mo-OH, these being the forms observed by e.p.r. The accepting group in the functional enzyme, for the proton transferred from the substrate while molybdenum is reduced in the catalytic reaction [Gutteridge, Tanner & Bray (1978) Biochem J. 175 869-878], is thus taken to be Mo=S.  相似文献   

4.
The mechanism of propagation of the radical center between the cofactor, substrate, and product in the adenosylcobalamin- (AdoCbl) dependent reaction of ethanolamine ammonia-lyase has been probed by pulsed electron nuclear double resonance (ENDOR) spectroscopy. The radical of S-2-aminopropanol, which appears in the steady state of the reaction, was used in ENDOR experiments to determine the nuclear spin transition frequencies of (2)H introduced from either deuterated substrate or deuterated coenzyme and of (13)C introduced into the ribosyl moiety of AdoCbl. A (2)H doublet (1.4 MHz splitting) was observed centered about the Larmor frequency of (2)H. Identical ENDOR frequencies were observed for (2)H irrespective of its mode of introduction into the complex. A (13)C doublet ENDOR signal was observed from samples prepared with [U-(13)C-ribosyl]-AdoCbl. The (13)C coupling tensor obtained from the ENDOR powder pattern shows that the (13)C has scalar as well as dipole-dipole coupling to the unpaired electron located at C1 of S-2-aminopropanol. The dipole-dipole coupling is consistent with a distance of 3.4+/-0.2 A between C1 of the radical and C5' of the labeled cofactor component. These results establish that the C5' carbon of the 5'-deoxyadenosyl radical moves approximately 7 A from its position as part of AdoCbl to a position where it is in contact with C1 of the substrate which lies approximately 12 A from the Co(2+) of cob(II)alamin. These findings are also consistent with the contention that 5'-deoxyadenosine is the sole mediator of hydrogen transfers in ethanolamine ammonia-lyase.  相似文献   

5.
The kinetics of the reduction of oxidized Fe-protein of nitrogenase from Azotobacter chroococcum by sodium dithionite were studied by stopped-flow and rapid-freezing e.p.r. (electron-paramagnetic-resonance) spectroscopy. The appearance of the gav. = 1.94 e.p.r. signal (0.24 electron integrated intensity/mol) was associated with a one-electron reduction by SO2--with k greater than 10(8)M-1-S-1 at 23 degrees C. A value of k = 1.75s-1 was obtained for the rate of dissociation of S2O42- into 2SO2-- at 23 degrees C. Further reductions by SO2-- occurred in three slower phases with rate constants in the range 10(4) -10(6)M-1-S-1. These latter phases have no corresponding e.p.r. signal changes and are probably associated with enzymically inactive protein. The high rate of reduction by SO2-- of the Fe-protein alone (k greater than 10(8)M-1-S-1) relative to the rate of oxidation of the Fe-protein in the catalytically active Fe:Mo-Fe protein complex (k = 2.2 X 1O(2)s-1) and the observation that in the steady state the Fe-protein is substantially oxidized means that at normal assay concentrations another reaction must limit the rate of reduction of Fe-protein during turnover.  相似文献   

6.
The kinetic mechanism of hog kidney D-amino acid oxidase with D-alpha-aminobutyrate as substrate has been examined in detail using a combination of steady state and rapid reaction methods. At concentrations of D-alpha-aminobutyrate below 0.5 mM, the rapid reaction and steady state results are consistent with the mechanism previously proposed for D-alanine (Massey, V., and Gibson, Q. H. (1964) Fed. Proc. 23, 18-29; Porter, D. J. T., Voet, J. G., and Bright, H. J. (1977) J. Biol. Chem. 252, 4464-4473). Both flavin reduction by D-alpha-aminobutyrate and reoxidation are quite rapid. Release of product from the oxidized enzyme has been measured directly and matches the turnover number at infinite concentrations of both substrates. Substitution of deuterium for the alpha-hydrogen decreases the rate of reduction 1.4-fold, without any effect on the apparent Kd. Computer simulations show that the kinetic isotope effects on the reductive half-reaction with D-alanine reported by Porter et al. (see above reference) can be explained using a two-step model with a kinetic isotope effect of 1.75 on the limiting rate of reduction. The effect of enzyme concentration on the kinetics has been examined in some detail. With D-alanine as substrate, increasing the enzyme concentration over the range 29 nM to 17 microM resulted in less than a 2-fold decrease in the turnover number. The Kd for benzoate binding also decreased marginally with increasing enzyme concentration. The effect of enzyme concentration is consistent with a decrease in the rate of release of ligands from the oxidized enzyme as the enzyme concentration is increased.  相似文献   

7.
The existence of rapid light-induced changes of light scattering in suspensions of bovine rod outer segment membranes has been described previously [H. Kühn et al. (1981) Proc. Natl Acad. Sci. USA, 78, 6873-6877]. The signal observed in the presence of GTP has been interpreted as being related to the rhodopsin-catalyzed exchange of GTP for GDP bound to the GTP-binding protein, i.e. to the formation of the activator of the cGMP phosphodiesterase [B.K.K. Fung et al. (1981) Proc. Natl Acad. Sci. USA, 78, 152-156]. We have tested this interpretation in the present paper by investigating the relation between the light-scattering signal and the activity of the phosphodiesterase using rapid recording techniques for both processes. All the results obtained are consistent with the above hypothesis. The amplitude of the light-scattering signal and the activity of the phosphodiesterase are shown to present the same dependence upon the flash intensity and upon the concentration of GTP or its analog guanosine 5'-[beta, gamma--imido]triphosphate (p[NH]ppG). The results suggest that the GTP-binding protein possesses one high-affinity p[NH]ppG-binding site (Kd much less than 0.1 microM). At high concentrations of GTP or p[NH]ppG the phosphodiesterase is activated in the dark and the light-scattering signal is correspondingly reduced; both effects are prevented by previous incubation with guanosine 5'-[beta-thio]diphosphate (p[S]pG).  相似文献   

8.
The oxidative half-reaction of phenol hydroxylase has been studied by stopped-flow spectrophotometry. Three flavin-oxygen intermediates can be detected when the substrate is thiophenol, or m-NH2, m-OH, m-CH3, m-Cl, or p-OH phenol. Intermediate I, the flavin C(4a)-hydroperoxide, has an absorbance maximum at 380-390 nm and an extinction coefficient approximately 10,000 M-1 cm-1. Intermediate III, the flavin C(4a)-hydroxide, has an absorbance maximum at 365-375 nm and an extinction coefficient approximately 10,000 M-1 cm-1. Intermediate II has absorbance maxima of 350-390 nm and extinction coefficients of 10,000-16,000 M-1 cm-1 depending on the substrate. A Hammett plot of the logarithm of the rates of the oxygen transfer step, the conversion of intermediate I to intermediate II, gives a straight line with a slope -0.5. Fluoride ion is a product of the enzymatic reaction when 2,3,5,6-tetrafluorophenol is the substrate. These results are consistent with an electrophilic substitution mechanism for oxygen transfer. The conversions of I to II and II to III are acid-catalyzed. A kinetic isotope effect of 8 was measured for the conversion of II to III using deuterated resorcinol as substrate. The conversion of III to oxidized enzyme is base-catalyzed, suggesting that the reaction depends on the removal of the flavin N(5) proton. Product release occurs at the same time as the formation of intermediate III, or rapidly thereafter. The results are interpreted according to the ring-opened model of Entsch et al. (Entsch, B., Ballou, D. P., and Massey, V. (1976) J. Biol. Chem. 251, 2550-2563).  相似文献   

9.
For isoquinoline 1-oxidoreductase (IsoOr), the reaction mechanism under turnover conditions was studied by EPR spectroscopy using rapid-freeze methods. IsoOr displays several EPR-active Mo(V) species including the "very rapid" component found also in xanthine oxidase (XanOx). For IsoOr, unlike XanOx or quinoline 2-oxidoreductase (QuinOr), this species is stable for about 1 h in the absence of an oxidizing substrate [Canne, C., Stephan, I., Finsterbusch, J., Lingens, F., Kappl, R., Fetzner, S., and Hüttermann, J. (1997) Biochemistry 36, 9780-9790]. Under rapid-freeze conditions in the presence of ferricyanide the very rapid species behaves as a kinetically competent intermediate present only during steady-state turnover. To explain the persistence of the very rapid species in IsoOr in the absence of an added oxidant, extremely slow product dissociation is required. This new finding that oxidative conditions facilitate decay of the very rapid signal for IsoOr supports the mechanism of substrate turnover proposed by Lowe, Richards, and Bray [Lowe, D. J., Richards, R. L., and Bray, R. C. (1997) Biochem. Soc. Trans. 25, 774-778]. Additional stopped-flow data reveal that alternative catalytic cycles occur in IsoOr and show that the product dissociates after transfer of a single oxidizing equivalent from ferricyanide. In rapid-freeze measurements magnetic interactions of the very rapid Mo(V) species and the iron-sulfur center FeSI of IsoOr and QuinOr were observed, proving that FeSI is located close to the molybdopterin cofactor in the two proteins. This finding is used to relate the two different iron-sulfur centers of the aldehyde oxidoreductase structure with the EPR-detectable FeS species of the enzymes.  相似文献   

10.
The hydrolysis of O-arylphosphorothioates by protein-tyrosine phosphatases (PTPases) was studied with the aim of providing a mechanistic framework for the reactions of this important class of substrate analogues. O-arylphosphorothioates are hydrolyzed 2 to 3 orders of magnitude slower than O-aryl phosphates by PTPases. This is in contrast to the solution reaction where phosphorothioates display 10-60-fold higher reactivity than the corresponding oxygen analogues. Kinetic analyses suggest that PTPases utilize the same active site and similar kinetic and chemical mechanisms for the hydrolysis of O-arylphosphorothioates and O-aryl phosphates. Thio substitution has no effect on the affinity of substrate or product for the PTPases. Bronsted analyses suggest that like the PTPase-catalyzed phosphoryl transfer reaction the transition state for the PTPase-catalyzed thiophosphoryl transfer is highly dissociative, similar to that of the corresponding solution reaction. The side chain of the active-site Arg residue forms a bidentate hydrogen bond with two of the terminal phosphate oxygens in the ground state and two of the equatorial oxygens in a transition state analog complex with vanadate [Denu et al. (1996) Proc. Natl. Acad. Sci. USA 93, 2493-2498; Zhang, M. et al. (1997) Biochemistry 36, 15-23; Pannifer et al. (1998) J. Biol. Chem. 273, 10454-10462]. Replacement of the active-site Arg409 in the Yersinia PTPase by a Lys reduces the thio effect by 54-fold, consistent with direct interaction and demonstrating strong energetic coupling between Arg409 and the phosphoryl oxygens in the transition state. These results suggest that the large thio effect observed in the PTPase reaction is the result of inability to achieve precise transition state complementarity in the enzyme active site with the larger sulfur substitution.  相似文献   

11.
Although X-ray crystallographic and NMR studies have been made on the adenylate kinases, the substrate-binding sites are not unequivocally established. In an attempt to shed light on the binding sites for MgATP2- and for AMP2- in human cytosolic adenylate kinase (EC 2.7.4.3, hAK1), we have investigated the enzymic effects of replacement of the arginine residues (R44, R132, R138, and R149), which had been assumed by Pai et al. [Pai, E. F., Sachsenheimer, W., Schirmer, R. H., & Schulz, G. E. (1977) J. Mol. Biol. 114, 37-45] to interact with the phosphoryl groups of AMP2- and MgATP2-. With use of the site-directed mutagenesis method, point mutations were made in the artificial gene for hAK1 [Kim, H. J., Nishikawa, S., Tanaka, T., Uesugi, S., Takenaka, H., Hamada, M., & Kuby, S. A. (1989) Protein Eng. 2, 379-386] to replace these arginine residues with alanyl residues and yield the mutants R44A hAK1, R132A hAK1, R138A hAK1, and R149A hAK1. The resulting large increases in the Km,app values for AMP2- of the mutant enzymes, the relatively small increases in the Km,app values for MgATP2-, and the fact that the R132A, R138A, and R149A mutant enzymes proved to be very poor catalysts are consistent with the idea that the assigned substrate binding sites of Pai et al. (1977) have been reversed and that their ATP-binding site may be assigned as the AMP site.  相似文献   

12.
P. S. Alban et al. (J. Appl. Microbiol. (1998) 85, 875-882) reported that a mutant H2O2-resistant strain of Spirullum (S.) volutans showed constitutive overexpression of a protein whose amino acid sequence and molecular weight closely resembled that of a subunit of rubrerythrin, a non-heme iron protein with no known function. They also reported that the mutant strain, but not the wild-type, showed NADH peroxidase activity. Here we demonstrate that rubrerythrin and nigerythrin from Desulfovibrio vulgaris and rubrerythrin from Clostridium perfringens show NADH peroxidase activities in an in vitro system containing NADH, hydrogen peroxide, and a bacterial NADH oxidoreductase. The peroxidase specific activities of the rubrerythrins with the "classical" heme peroxidase substrate, o-dianisidine, are many orders of magnitude lower than that of horseradish peroxidase. These results are consistent with the phenotype of the H2O2-resistant strain of S. volutans. The reaction of reduced (i.e., all-ferrous) rubrerythrin with excess O2 takes several minutes, whereas the anaerobic reaction of reduced rubrerythrin with hydrogen peroxide is on the millisecond time scale and results in full oxidation of all iron centers to their ferric states. Rubrerythrins could, thus, function as the terminal components of NADH peroxidases in air-sensitive bacteria and archaea.  相似文献   

13.
Xanthine oxidase is stable and active in aqueous dimethyl sulphoxide solutions of up to at least 57% (w/w). Simple techniques are described for mixing the enzyme in this solvent at--82 degrees C, with its substrate, xanthine. When working at high pH values under such conditions, no reaction occurred, as judged by the absence of e.p.r. signals. On warming to--60 degrees C, for 10 min, however, the Very Rapid molybdenum(V) e.p.r. signal was obtained. This signal did not change on decreasing the pH, while maintaining the sample in liquid nitrate reductase, caused its molybdenum(V) e.p.r. signal to change from the high-pH to the low-pH form. These findings are not compatible with the conclusions of Edmondson, Ballou, Van Heuvelen, Palmer & Massey [J. Biol. Chem. (1973) 248, 6135-6144], that the Very Rapid signal is in prototropic equilibrium with the Rapid signal, and should be important in understanding the mechanism of action of the enzyme. They emphasize the unique nature of the intermediate represented by the Very Rapid e.p.r. signal. The possible value of the pK for loss of an exchangeable proton from the Rapid signal is discussed.  相似文献   

14.
Silverman et al. (1979. J. Am. Chem. Soc. 101:6734-6740) have reported measurements of the loss of 18O to solvent from the isotopically labeled CO2--HCO3-system and of the mixing of 18O and 13C labels within the system, as catalyzed by human carbonic anhydrase C in the pH range 6-8. This work is an extension of earlier work (Silverman and Tu. 1976. J. Am. Chem. Soc. 98:978-984) on the very similar bovine enzyme. The more recent work is analyzed by its authors in terms of the "hydroxide" model for the apparent pH-dependence of enzymatic activity, a model in which the pH-dependence is associated with the presumed ionization of an H2O ligand of the active-site metal ion to OH-. From a comparison of their data with a solution of the coupled differential equations that describe the kinetics of isotope exchange in terms of the model, Silverman et al. derived a pH-dependent rate of exchange for the water molecule which is formed at the active site of the enzyme during dehydration. By contrast, using the same data and a model in which active enzyme has a water molecule on the metal ion at the active site, and similar differential equations, we derive a value for the rate of exchange of water that is pH-independent. This model has the attraction that it explains the magnetic relaxation rate of solvent water protons in the Co2+-substituted enzyme, whereas the hydroxide mechanism cannot explain these data without the introduction of unfounded ad hoc assumptions; further, the presence of an OH- ligand of the metal has never been demonstrated. We also include an analysis of analogous data for the bovine enzyme. One result of our analysis is that the pKa for activity of the enzyme samples used is near 6.0, implying that the bulk of the data were taken when the enzyme was essentially all active. It is straightforward to account for the pH-dependence of the data near and below the pKa by using an empirically-derived value for the pKa. However, we have recently developed a model for the low pH (inactive) enzyme that has been successful in interpreting a wide range of data, and we show that this new view can explain the few points at low pH quite adequately. Additionally, we consider the recent kinetic results for the human C enzyme, obtained at chemical equilibrium by studies of the linewidths of nuclear magnetic resonances of 13C in labeled substrate (Simonsson et al. 1979. Eur. J. Biochem. 93:409-417) and show that these experiments and those of Silverman et al. are all consistent with kinetic data from nonequilibrium stopped-flow experiments, viewed in terms of our model, in the limit of low substrate concentration. Results at higher concentrations indicate that the Michaelis constants and equilibrium constants differ somewhat.  相似文献   

15.
Marciniszyn et al. (Marciniszyn, J., Huang, J. S. Hartsuch, J. A., Tang, J. (1976) J. Biol. Chem. 251, 7095-7102) have recently suggested an intermediate in the intramolecular activation of pepsinogen. As evidence, they showed apparent competitive inhibition of activation by globin, indication a pepsinogen-globin complex. Previous work had shown pepsinogen activation to occur very rapidly in the presence of high concentrations of hemoglobin, a very similar pepsin substrate (McPhie, P. (1974) Biochem. Biophys. Res. Commun. 56, 789-792). This contradiction has been resolved by a re-evaluation of the techniques used in the two investigations. The experimental conditions of Marciniszyn et al. Were inadequately defined to ensure denaturation of pepsin, a prerequisite of their method. A small decrease in pH, caused by the presence of extraneous protein, prevents this denaturation and leads to consistent underestimates of the rate of zymogen activation.  相似文献   

16.
The reaction of [nitrotyrosyl]cytochrome c with ligands was studied by stopped-flow techniques. At pH 7.0 the reaction with imidazole shows two distinct phases, one fast phase being concentration-dependent and a slow phase being concentration-independent. The results are consistent with the existence of two forms of [nitrotyrosyl]cytochrome c in solutions [Schejter et al. (1970) Biochemistry 9, 5118-5122]; form I, the smaller fraction, seems to be responsible for the slow first-order process.  相似文献   

17.
Cytochrome P-450-catalyzed dehydrogenation of 1,4-dihydropyridines   总被引:1,自引:0,他引:1  
A variety of different 4-substituted 1,4-dihydropyridine Hantzsch esters are substrates for ring dehydrogenation by a cytochrome P-450 (P-450) enzyme (P-450 UT-A); the substitutent could be varied from a hydrogen to a naphthalenyl, but a pyrenyl derivative was not dehydrogenated. When a 4-alkyl group is present, both the P-450 which oxidizes the substrate and other P-450s can be inactivated (by putative alkyl radicals). P-450s did not discriminate with regard to removal of the 4-H atoms from an enantiomeric pair of dihydropyridines. Losses of the 4-proton and N-methyl from a N-methyl-1,4-dihydropyridine occur at similar rates. The calculated intrinsic kinetic hydrogen isotope effect (Dk) for dehydrogenation of 1,4-dihydro-2,6-dimethyl-4-phenyl-3,5-pyridinedicarboxylic acid dimethyl ester was 2.9 in a reconstituted P-450 UT-A enzyme system. No significant kinetic hydrogen isotope effect was observed in microsomal incubations for the dehydrogenation of this compound or 1,4-dihydro-2,6-dimethyl-3,5-pyridinedicarboxylic acid diethyl ester in a variety of competitive and noncompetitive experiments. In light of previous studies on the magnitude of kinetic hydrogen isotope effects in P-450 systems (e.g. Miwa et al., 1983 (Miwa, G. T., Walsh, J. S., Kedderis, G. L., and Hollenberg, P. F. (1983) J. Biol. Chem. 258, 14445-14449], the mechanistic proposals of Augusto et al., 1982 (Augusto, O., Beilan, H. S., and Ortiz de Montellano, P. R. (1982) J. Biol. Chem. 257, 11288-11295)) for enzyme inactivation by 4-alkyl-substituted Hantzsch pyridine esters, and other precedents for sequential electron transfer in amine oxidation by P-450s, we interpret these results as being consistent with P-450-mediated 1-electron oxidation of dihydropyridines followed by the facile loss of the 4-proton, with subsequent electron transfer to complete the reaction.  相似文献   

18.
Studies by e.p.r. (electron-paramagnetic-resonance) spectroscopy and by stopped-flow spectrophotometry on turkey liver xanthine dehydrogenase revealed strong similarities to as well as important differences from the Veillonella alcalescens xanthine dehydrogenase and milk xanthine oxidase. The turkey enzyme is contaminated by up to three non-functional forms, giving molybdenum e.p.r. signals designated Resting I, Resting II and Slow. Slow and to a lesser extent Resting I signals are like those from the Veillonella enzyme, whereas Resting II is very like a resting signal described by K. V. Rajagopolan, P. Handler, G. Palmer & H. Beinert (1968) (J. Biol. Chem. 243, 3784-3796) for aldehyde oxidase. Another non-functional form that gives the Inhibited signal is produced on treatment of the enzyme with formaldehyde. Stopped-flow measurements at 450 nm show that, as for the milk enzyme, reduction by xanthine is rate-limiting in enzyme turnover. The active enzyme gives rise to Very Rapid and Rapid molybdenum(V) e.p.r. signals, as well as to an FADH signal. That these signals are almost indistinguishable from those of the milk enzyme, confirms the similarities between the active sites. There are two types of iron-sulphur centres that give signals like those in the milk enzyme, though with slightly different parameters. Quantitative reduction titration of the functional enzyme with xanthine revealed two important differences between the turkey and the milk enzymes. First, the turkey enzyme FADH/FADH2 system has a redox potential sufficiently low that xanthine is incapable of reducing the flavin completely. This finding presumably explains the very low oxidase activity. Secondly, whereas the Fe/S II chromophore in the milk enzyme has a relatively high redox potential, for the turkey enzyme the value of this potential is lower and similar to that of its Fe/S I chromophore.  相似文献   

19.
Ribbons et al. (Ribbons, D.W., Ohta, Y., and Higgins, I.J. (1972) in Molecular Basis of Electron Transport, Miami Winter Symposic Series (Schultz, J., and Cameron, B.F., eds) Vol. 4, pp. 251-274, Academic Press, New York) presented a preliminary report that the flavoenzyme monooxygenase orcinol hydroxylase shows mixed type 4R, 4S stereospecificity with respect to dihydronicotinamide oxidation when resorcinol and m-cresol were used as substrate analogs. With the natural substrate orcinol, 4R chirality was maintained. In kinetic isotope experiments reported here, we demonstrate in fact that orcinol hydroxylase maintains 4R stereospecificity with respect to dihydronicotinamide oxidation with all three substrates, orcinol, resorcinol, and m-cresol. Deuterium and tritium kinetic isotope effects were detected under Vmax conditions with (4R)-[4-2H]-, and (4R)-[4-3H]NADH for all three substrates. No isotope effect was observed with (4S)-[4-2H]NADH and tritium labilization from assays with (4S)-[4-3H]-NADH was negligible in all cases.  相似文献   

20.
The P3HR-1 subclone of Jijoye differs from Jijoye and from other Epstein-Barr virus (EBV)-infected cell lines in that the virus produced by P3HR-1 cultures lacks the ability to growth-transform normal B lymphocytes (Heston et al., Nature (London) 295:160-163, 1982; Miller et al., J. Virol. 18:1071-1080, 1976; Miller et al., Proc. Natl. Acad. Sci. U.S.A. 71:4006-4010, 1974; Ragona et al., Virology 101:553-557, 1980). The P3HR-1 virus was known to be deleted for a region which encodes RNA in latently infected, growth-transformed cells (Bornkamm et al., J. Virol. 35:603-618, 1980; Heller et al., J. Virol. 38:632-648, 1981; King et al., J. Virol. 36:506-518, 1980; Raab-Traub et al., J. Virol. 27:388-398, 1978; van Santen et al., Proc. Natl. Acad. Sci. U.S.A. 78:1930-1934, 1980). This deletion is now more precisely defined. The P3HR-1 genome contains less than 170 base pairs (and possibly none) of the 3,300-base pair U2 region of EBV DNA and is also lacking IR2 (a 123-base pair repeat which is the right boundary of U2). A surprising finding is that EBV isolates vary in part of the U2 region. Two transforming EB viruses, AG876 and Jijoye, are deleted for part of the U2 region including most or all of a fragment, HinfI-c, which encodes part of one of the three more abundant cytoplasmic polyadenylated RNAs of growth-transformed cells (King et al., J. Virol. 36:506-518, 1980; King et al., J. Virol. 38:649-660, 1981; van Santen et al., Proc. Natl. Acad. Sci. U.S.A. 78:1930-1934).  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