首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Chemical modification of phenol hydroxylase by ethoxyformic anhydride   总被引:1,自引:0,他引:1  
Phenol hydroxylase was inactivated by ethoxyformic anhydride. Part of the inactivation was related to modification of histidyl residues. The remaining part of the inactivation is proposed to be due to the modification of a lysyl residue which, we suggest, is identical with the one previously described, being essential for the binding of NADPH [Neujahr, H. Y. and Kjellén, K. G. (1980) Biochemistry 19, 4967-4972]. The overall inactivation reaction is biphasic and follows pseudo-first-order kinetics. Numerical analysis of kinetic data was applied to discriminate between simultaneous reactions at different sites. It is proposed that phenol hydroxylase contains two essential histidyl residues, located in or near the NADPH-binding sites. Ethoxyformylation of the lysyl residue(s) caused tightening of the binding of phenol and perturbation of the FAD spectrum of phenol hydroxylase, similar to that caused by phenolic effectors.  相似文献   

2.
Phenol hydroxylase (PH) from Acinetobacter radioresistens S13 represents an example of multicomponent aromatic ring monooxygenase made up of three moieties: a reductase (PHR), an oxygenase (PHO) and a regulative component (PHI). The function of the oxygenase component (PHO), here characterized for the first time, is to bind molecular oxygen and catalyse the mono-hydroxylation of substrates (phenol, and with less efficiency, chloro- and methyl-phenol and naphthol). PHO was purified from extracts of A. radioresistens S13 cells and shown to be a dimer of 206 kDa. Each monomer is composed by three subunits: alpha (54 kDa), beta (38 kDa) and gamma (11 kDa). The gene encoding PHO alpha (named mopN) was cloned and sequenced and the corresponding amino acid sequence matched with that of functionally related oxygenases. By structural alignment with the catalytic subunits of methane monooxygenase (MMO) and alkene monooxygenase, we propose that PHO alpha contains the enzyme active site, harbouring a dinuclear iron centre Fe-O-Fe, as also suggested by spectral analysis. Conserved hydrophobic amino acids known to define the substrate recognition pocket, are also present in the alpha-subunit. The prevalence of alpha-helices (99.6%) as studied by CD confirmed the hypothized structural homologies between PHO and MMO. Three parameters (optimum ionic strength, temperature and pH) that affect kinetics of the overall phenol hydroxylase reaction were further analyzed with a fixed optimal PHR/PHI/PHO ratio of 2/1/1. The highest level of activity was evaluated between 0.075 and 0.1 m of ionic strength, the temperature dependence showed a maximum of activity at 24 degrees C and finally the pH for optimal activity was determined to be 7.5.  相似文献   

3.
This paper reports the isolation and characterization of the regulatory moiety of the multicomponent enzyme phenol hydroxylase from Acinetobacter radioresistens S13 grown on phenol as the only carbon and energy source. The whole enzyme comprises an oxygenase moiety (PHO), a reductase moiety (PHR) and a regulatory moiety (PHI). PHR contains one FAD and one iron-sulfur cluster, whose function is electron transfer from NADH to the dinuclear iron centre of the oxygenase. PHI is required for catalysis of the conversion of phenol to catechol in vitro, but is not required for PHR activity towards alternative electron acceptors such as cytochrome c and Nitro Blue Tetrazolium. The molecular mass of PHI was determined to be 10 kDa by SDS/PAGE, 8.8 kDa by MALDI-TOF spectrometry and 18 kDa by gel-permeation. This finding suggests that the protein in its native state is a homodimer. The isoelectric point is 4.1. PHI does not contain any redox cofactor and does not bind ANS, a fluorescent probe for hydrophobic sites. The N-terminal sequence is similar to those of the regulatory proteins of phenol hydroxylase from A. calcoaceticus and Pseudomonas CF 600. In the reconstituted system, optimal reaction rate was achieved when the stoichiometry of the components was 2 PHR monomers: 1 PHI dimer: 1 PHO (alphabetagamma) dimer. PHI interacts specifically with PHR, promoting the enhancement of FAD fluorescence emission. This signal is diagnostic of a conformational change of PHR that might result in a better alignment with respect to PHO.  相似文献   

4.
Circular dichroism spectra and circular dichroism difference spectra, generated when porcine heart mitochondrial and supernatant malate dehydrogenase bind coenzymes or when enzyme dihydroincotinamide nucleotide binary complexes bind substrate analogs, are presented. No significant changes are observed in protein chromophores in the 200- to 240-nm spectral range indicating that there is apparently little or no perturbation of the alpha helix or peptide backbone when binary or ternary complexes are formed. Quite different spectral perturbances occur in the two enzymes with reduced coenzyme binding as well as with substrate-analog binding by enzyme-reduced coenzyme binding. Comparison of spectral perturbations in both enzymes with oxidized or reduced coenzyme binding suggests that the dihydronicotinamide moiety of the coenzyme interacts with or perturbs indirectly the environment of aromatic amino acid residues. Reduced coenzyme binding apparently perturbs tyrosine residues in both mitochondrial malate dehydrogenase and lactic dehydrogenase. Reduced coenzyme binding perturbs tyrosine and tryptophan residues in supernatant malate dehydrogenase. The number of reduced coenzyme binding sites was determined to be two per 70,000 daltons in the mitochondrial enzyme, and the reduced coenzyme dissociation constants, determined through the change in ellipticity at 260 nm, with dihydronicotinamide adenine dinucleotide binding, were found to be good agreement with published values (Holbrook, J. J., and Wolfe, R. G. (1972) Biochemistry 11, 2499-2502) obtained through fluorescence-binding studies and indicate no apparent extra coenzyme binding sites. When D-malate forms a ternary complex with malate dehydrogenase-reduced coenzyme complexes, perturbation of both adenine and dihydronicotinamide chromophores is evident. L-Malate binding, however, apparently produces only a perturbation of the adenine chromophore in such complexes. Since the coenzyme has been found to bind in an open conformation on the surface of the enzyme and the substrate analogs bind at or very near the dihydronicotinamide moiety binding site, protein conformational changes are implicated during ternary complex formation with D-malate which can effect the adenine chromophore at some distance from the substrate binding site.  相似文献   

5.
W H Ward  A R Fersht 《Biochemistry》1988,27(15):5525-5530
Tyrosyl-tRNA synthetase from Bacillus stearothermophilus is a classical example of an enzyme with half-of-the-sites activity. The enzyme crystallizes as a symmetrical dimer that is composed of identical subunits, each having a complete active site. In solution, however, tyrosyl-tRNA synthetase binds tightly, and activates rapidly, only 1 mol of Tyr/mol of dimer. It has recently been shown that the half-of-the-sites activity results from an inherent asymmetry of the enzyme. Only one subunit catalyzes formation of Tyr-AMP, and interchange of activity between subunits is not detectable over a long time scale. Paradoxically, however, the kinetics of tRNA charging are biphasic with respect to [Tyr], suggesting that both subunits of the dimer are catalytically active. This paradox has now been resolved by kinetic analysis of heterodimeric enzymes containing different mutations in each subunit. Biphasic kinetics with unchanged values of KM for Tyr are maintained when one of the two tRNA-binding domains is removed and also when the affinity of the "inactive" site for Try is reduced by 2-58-fold. The biphasic kinetics do not result from catalysis at both active sites, but instead appear to result from two molecules of Tyr binding sequentially to the same site. A second molecule of Tyr perhaps aids the dissociation of Tyr-tRNA by displacing the tyrosyl moiety from its binding site. A monomer of the enzyme is probably too small to allow both recognition and aminoacylation of a tRNA molecule. This could explain the requirement for the enzyme to function as an asymmetric dimer.  相似文献   

6.
D W Pettigrew  G J Yu  Y Liu 《Biochemistry》1990,29(37):8620-8627
Substrate binding to Escherichia coli glycerol kinase (EC 2.7.1.30; ATP-glycerol 3-phosphotransferase) was investigated by using both kinetics and binding methods. Initial-velocity studies in both reaction directions show a sequential kinetic mechanism with apparent substrate activation by ATP and substrate inhibition by ADP. In addition, the Michaelis constants differ greatly from the substrate dissociation constants. Results of product inhibition studies and dead-end inhibition studies using 5'-adenylyl imidodiphosphate show the enzyme has a random kinetic mechanism, which is consistent with the observed formation of binary complexes with all the substrates and the glycerol-independent MgATPase activity of the enzyme. Dissociation constants for substrate binding determined by using ligand protection from inactivation by N-ethylmaleimide agree with those estimated from the initial-velocity studies. Determinations of substrate binding stoichiometry by equilibrium dialysis show half-of-the-sites binding for ATP, ADP, and glycerol. Thus, the regulation by nucleotides does not appear to reflect binding at a separate regulatory site. The random kinetic mechanism obviates the need to postulate such a site to explain the formation of binary complexes with the nucleotides. The observed stoichiometry is consistent with a model for the nucleotide regulatory behavior in which the dimer is the enzyme form present in the assay and its subunits display different substrate binding affinities. Several properties of the enzyme are consistent with negative cooperativity as the basis for the difference in affinities. The possible physiological importance of the regulatory behavior with respect to ATP is considered.  相似文献   

7.
H Y Neujahr 《Biochemistry》1988,27(10):3770-3775
Spectrophotometric titration of phenol hydroxylase (EC 1.14.13.7) with phenol indicated interacting sites for phenol binding. In the absence of added thiol, the cooperativity was positive up to a pH around 8.0 but negative at higher pH values. With added thiol-ethylenediaminetetraacetate, the cooperativity was negative at all investigated pH values. Conversely, a corresponding titration of an enzyme preparation that had been selectively modified in its two most reactive SH groups indicated positive cooperativity at all studied pH values. This selective modification affects the activity of the enzyme to a very minor degree, in contrast to more extensive SH blocking, which displaces flavin adenine dinucleotide with a corresponding loss of activity [Neujahr, H. Y., & Gaal, A. (1975) Eur. J. Biochem. 58, 351-357]. The reactivity of SH groups in the enzyme was significantly decreased after turnover. Thiol treatment restored it to that of the native enzyme. Adding phenol prior to reduced nicotinamide adenine dinucleotide phosphate (NADPH) in the assay of phenol hydroxylase gave immediate linearity and higher initial rates than when NADPH was added first. In the absence of added thiol, there was then a shift of the pH optimum. The results indicate slow conformational changes limiting the rate of the overall reaction. The two most reactive SH groups of phenol hydroxylase, though not participating in any obvious redox reactions, are important for these slow conformational changes and for the cooperativity of phenol-binding sites, wherein the anionic S- forms may be involved (pKa for cysteine is 8.35).  相似文献   

8.
Angucyclines are aromatic polyketides produced in Streptomycetes via complex enzymatic biosynthetic pathways. PgaE and CabE from S. sp PGA64 and S. sp. H021 are two related homo-dimeric FAD and NADPH dependent aromatic hydroxylases involved in the early steps of the angucycline core modification. Here we report the three-dimensional structures of these two enzymes determined by X-ray crystallography using multiple anomalous diffraction and molecular replacement, respectively, to resolutions of 1.8 A and 2.7 A. The enzyme subunits are built up of three domains, a FAD binding domain, a domain involved in substrate binding and a C-terminal thioredoxin-like domain of unknown function. The structure analysis identifies PgaE and CabE as members of the para-hydroxybenzoate hydroxylase (pHBH) fold family of aromatic hydroxylases. In contrast to phenol hydroxylase and 3-hydroxybenzoate hydroxylase that utilize the C-terminal domain for dimer formation, this domain is not part of the subunit-subunit interface in PgaE and CabE. Instead, dimer assembly occurs through interactions of their FAD binding domains. FAD is bound non-covalently in the "in"-conformation. The active sites in the two enzymes differ significantly from those of other aromatic hydroxylases. The volumes of the active site are significantly larger, as expected in view of the voluminous tetracyclic angucycline substrates. The structures further suggest that substrate binding and catalysis may involve dynamic rearrangements of the middle domain relative to the other two domains. Site-directed mutagenesis studies of putative catalytic groups in the active site of PgaE argue against enzyme-catalyzed substrate deprotonation as a step in catalysis. This is in contrast to pHBH, where deprotonation/protonation of the substrate has been suggested as an essential part of the enzymatic mechanism.  相似文献   

9.
The phenol-degrading strain Trichosporon cutaneum R57 utilizes various aromatic and aliphatic compounds as a sole carbon and energy source. The intracellular activities of phenol hydroxylase [EC 1.14.13.7] of a Trichosporon cutaneum R57 strain grown on phenol (0.5 g/l) were measured. Different toxic phenol derivatives (cresols, nitrophenols and hydroxyphenols) were used as substrates in the reaction mixture for determination of the enzyme activity. The data obtained showed that the investigated enzyme was capable to hydroxylate all applied aromatic substrates. The measured activities of phenol hydroxylase varied significantly depending on the aromatic compounds used as substrates. The rate of phenol hydroxylase activity with phenol as a substrate (1.0 U/mg total cell protein) was accepted as 100%.  相似文献   

10.
M F Carlier  D Pantaloni 《Biochemistry》1976,15(21):4703-4712
The binding of reduced nicotinamide adenine dinucleotide phosphate (NADPH) to nicotinamide adenine dinucleotide phosphate (NADP) dependent isocitrate dehydrogenase from beef liver cytoplasm was studied by several equilibrium techniques (ultracentrifugation, molecular sieving, ultrafiltration, fluorescence). Two binding sites (per dimeric enzyme molecule) were found with slightly different dissociation constants (0.5 and 0.12 muM) and fluorescence yields (7.7 and 6.3). A ternary complex was formed between enzyme, isocitrate, and NADPH, in which NADPH dissociation constant was 5 muM. On the contrary, no binding of NADPH to the enzyme took place in the presence of magnesium isocitrate. Dialysis experiments showed the existence of 1 NADP binding site/dimer, with a dissociation constant of 26 muM. When NADPH was present with the enzyme in the proportion of 1 molecule/dimer, the dissociation constant of NADP was decreased fourfold, reaching a value quantitatively comparable to the Michaelis constant. The kinetics of coenzyme binding was followed using the stopped-flow technique with fluorescence detection. NADPH binding to the enzyme occurred through one fast reaction (k1 = 20 muM-1 s-1). Dissociation of NADPH took place upon NADP binding; however, equilibrium as well as kinetic data were incompatible with a simple competition scheme. Dissociation of NADPH from the enzyme upon magnesium isocitrate binding was preceded by the formation of a transitory ternary complex in which the fluorescence of NADPH was only about 30% of that in the enzyme-NADPH complex. Then interaction between the conenzymes and the involvement of ternary complexes in the catalytic mechanism are discussed in relation with what is known about the regulatory role of the coenzyme (Carlier, M. F., and Pantaloni, D. (1976), Biochemistry, 15, 1761-1766).  相似文献   

11.
The hydroxylation of phenylalanine to tyrosine by the liver enzyme phenylalanine hydroxylase is regulated by the level of phenylalanine. Whether there is a distinct allosteric binding site for phenylalanine outside of the active site has been unclear. The enzyme contains an N-terminal regulatory domain that extends through Thr117. The regulatory domain of rat phenylalanine hydroxylase was expressed in Escherichia coli. The purified protein behaves as a dimer on a gel filtration column. In the presence of phenylalanine, the protein elutes earlier from the column, consistent with a conformational change in the presence of the amino acid. No change in elution is seen in the presence of the non-activating amino acid proline. 1H–15N HSQC NMR spectra were obtained of the 15N-labeled protein alone and in the presence of phenylalanine or proline. A subset of the peaks in the spectrum exhibits chemical shift perturbation in the presence of phenylalanine, consistent with binding of phenylalanine at a specific site. No change in the NMR spectrum is seen in the presence of proline. These results establish that the regulatory domain of phenylalanine hydroxylase can bind phenylalanine, consistent with the presence of an allosteric site for the amino acid.  相似文献   

12.
The activation of yeast enolase by cobaltous ion in 0.1 M KCl is characterized by an activation constant of 1 microM and an inhibition constant of 18 microM. Measurements of binding of Co2+ to the apoenzyme show that a maximum of four Co2+ ions are bound per dimer in the presence or absence of substrate although binding is far tighter in the presence of substrate. Ultraviolet spectral titrations show evidence for a conformational change due exclusively to the binding of the first two ions of Co2+. Both visible and EPR spectra confirm that the environment of the first pair of cobalt ions ("conformational sites") is markedly different from that of the second pair in the "catalytic" sites. Cobalt at the conformational site appears to be a tetragonally distorted octahedral complex while the second pair of metal ions appears to be in a more regular tetrahedral symmetry. Addition of either Mg2+ or substrate to the enzyme with only one pair of cobalt ions per dimer causes striking changes in the metal ion environment. The conformational metal sites appear sufficiently shielded from solvent to be inaccessible to oxidation by H2O2, in contrast to the second pair of cobaltous ions whose ready oxidation by H2O2 inactivates the enzyme. Comparison of kinetic and binding data suggests that only one site of the dimeric enzyme can be active, since activity requires more than two metals bound per dimer and inactivation results from the binding of the fourth ion per dimer.  相似文献   

13.
The transient kinetics of aldehyde reduction by NADH catalyzed by liver alcohol dehydrogenase consist of two kinetic processes. This biphasic rate behavior is consistent with a model in which one of the two identical subunits in the enzyme is inactive during the reaction at the adjacent protomer. Alternatively, enzyme heterogeneity could result in such biphasic behavior. We have prepared liver alcohol dehydrogenase containing a single major isozyme; and the transient kinetics of this purified enzyme are biphasic.Addition of two [14C]carboxymethyl groups per dimer to the two “reactive” sulfhydryl groups (Cys46) yields enzyme which is catalytically inactive toward alcohol oxidation. Alkylated enzyme, as initially isolated by gel filtration chromatography at pH 7·5, forms an NAD+-pyrazole complex. However, the ability to bind NAD+-pyrazole is rapidly lost in pH 8·75 buffer; therefore, our alkylated preparations, as isolated by chromatography at pH 8·75, are inactive toward NAD+-pyrazole complex formation. We have prepared partially inactivated enzyme by allowing iodoacetic acid to react with liver alcohol dehydrogenase until 50% of the NAD+-pyrazole binding capacity remains; under these reaction conditions one [14C]carboxymethyl group is added per dimer. This partially alkylated enzyme preparation is isolated by gel filtration and has been aged sufficiently to lose NAD+-pyrazole binding ability at alkylated subunits. When solutions of native liver alcohol dehydrogenase and partially alkylated liver alcohol dehydrogenase containing the same number of unmodified active sites are allowed to react with substrate under single turnover conditions, partially alkylated enzyme is only half as reactive as native enzyme. This indicates that some molecular species in partially alkylated liver alcohol dehydrogenase that react with pyrazole and NAD+ during the active site titration do not react with substrate. These data are consistent with a model in which a subunit adjacent to an alkylated protomer in the dimeric enzyme is inactive toward substrate. In addition, NAD+-pyrazole binding at the protomers adjacent to alkylated subunits is slowly lost so that 75% of the enzyme-NAD+-pyrazole binding capacity is lost in 50% alkylated enzyme. These data supply strong evidence for subunit interactions in liver alcohol dehydrogenase.Binding experiments performed on partially alkylated liver alcohol dehydrogenase indicate that coenzyme binding is normal at a subunit adjacent to an alkylated protomer even though active ternary complexes cannot be formed. One hypothesis consistent with these results is the unavailability of zinc for substrate binding at the active site in subunits adjacent to alkylated protomers in monoalkylated dimer.  相似文献   

14.
Candida tropicalis isolated from acclimated activated sludge was used in this study. Cell suspensions with 5 x 10(7) cells ml(-1) were irradiated by using a He-Ne laser. After mutagenesis, the irradiated cell suspension was diluted and plated on yeast extract-peptone-dextrose (YEPD) medium. Plates with approximately 20 individual colonies were selected, and all individual colonies were harvested for phenol biodegradation. The phenol biodegradation stabilities for 70 phenol biodegradation-positive mutants, mutant strains CTM 1 to 70, ranked according to their original phenol biodegradation potentials, were tested continuously during transfers. Finally, mutant strain CTM 2, which degraded 2,600 mg liter(-1) phenol within 70.5 h, was obtained on the basis of its capacity and hereditary stability for phenol biodegradation. The phenol hydroxylase gene sequences were cloned in wild and mutant strains. The results showed that four amino acids were mutated by irradiation with a laser. In order to compare the activity of phenol hydroxylase in wild and mutant strains, their genes were expressed in Escherichia coli BL21(DE3) and enzyme activities were spectrophotometrically determined. It was clear that the activity of phenol hydroxylase was promoted after irradiation with a He-Ne laser. In addition, the cell growth and intrinsic phenol biodegradation kinetics of mutant strain CTM 2 in batch cultures were also described by Haldane's kinetic equation with a wide range of initial phenol concentrations from 0 to 2,600 mg liter(-1). The specific growth and degradation rates further demonstrated that the CTM 2 mutant strain possessed a higher capacity to resist phenol toxicity than wild C. tropicalis did.  相似文献   

15.
p-Hydroxybenzoate hydroxylase (EC 1.14.13.2) from Pseudomonas fluorescens is a NADPH-dependent, FAD-containing monooxygenase catalyzing the hydroxylation of p-hydroxybenzoate to form 3,4-dihydroxybenzoate in the presence of NADPH and molecular oxygen. The mechanism of this three-substrate reaction was investigated in detail at pH 6.6, 4 degrees C, by steady state kinetics, stopped flow spectrophotometry, and equilibrium binding experiments. The initial velocity patterns are consistent with a ping-pong type mechanism which involves two ternary complexes between the enzyme and substrates. The first ternary complex is formed by random addition of p-hydroxybenzoate and NADPH to the enzyme, followed by the release of the first product (NADP+). The reduced enzyme . p-hydroxybenzoate complex now reacts with oxygen, the third substrate, to form the second ternary complex. The enzyme-bound p-hydroxybenzoate then reacts with the activated oxygen to give 3,4-dihydroxybenzoate which is released regenerating the oxidized enzyme for the next cycle. The binding of p-hydroxybenzoate to the oxidized enzyme to form a 1:1 complex causes large, characteristic spectral perturbations and fluorescence quenching. The dissociation constant for the enzyme . substrate complex was obtained by titrations in which absorbance and/or fluorescence quenching was measured. The binding constants of NADPH to the enzyme with and without p-hydroxybenzoate were determined kinetically by measuring the rate of reduction of the enzyme at different concentrations of NADPH. The reduction of the enzyme proceeds extremely slowly in the absence of p-hydroxybenzoate. The presence of the substrate causes a dramatic stimulation (140,000-fold) in the rate of enzyme reduction. The anaerobic reduction of the enzyme by NADPH in the presence of p-hydroxybenzoate produces a transient charge-transfer intermediate. On the basis of the proposed mechanism, the dissociation constants for p-hydroxybenzoate and NADPH as well as the Michaelis constants for all the three substrates were calculated from the initial velocity data. The agreement obtained between various kinetic parameters from the initial rate measurements and those calculated from the individual rate constants determined in rapid reactions, strongly supports the proposed mechanism for the p-hydroxybenzoate hydroxylase reaction.  相似文献   

16.
We present here X-ray scattering data that yield new structural information on the multicomponent enzyme methane monooxygenase and its components: a hydroxylase dimer, and two copies each of a reductase and regulatory protein B. Upon formation of the enzyme complex, the hydroxylase undergoes a dramatic conformational change that is observed in the scattering data as a fundamental change in shape of the scattering particle such that one dimension is narrowed (by 25% or 24 A) while the longest dimension increases (by 20% or 25 A). These changes also are reflected in a 13% increase in radius of gyration upon complex formation. Both the reductase and protein B are required for inducing the conformational change. We have modeled the scattering data for the complex by systematically modifying the crystal structure of the hydroxylase and using ellipsoids to represent the reductase and protein B components. Our model indicates that protein B plays a role in optimizing the interaction between the active centers of the reductase and hydroxylase components, thus, facilitating electron transfer between them. In addition, the model suggests reasons why the hydroxylase exists as a dimer and that a possible role for the outlying gamma-subunit may be to stabilize the complex through its interaction with the other components. We further show that proteolysis of protein B to form the inactive B' results in a conformational change and B' does not bind to the hydroxylase. The truncation thus could represent a regulatory mechanism for controlling the enzyme activity.  相似文献   

17.
Mixing ATP with soluble oxidized cytochrome c oxidase induces a spectral perturbation in the Soret region of the enzyme. This spectral perturbation is observed at ATP concentrations similar to those found to modulate the catalytic activity of cytochrome c oxidase [Malatesta, Antonini, Sarti & Brunori (1987) Biochem. J. 248, 161-165]. The process is reversible and corresponds to a simple binding with Kd = 0.2 mM at 25 degrees C. The absorbance change follows a first-order time course, and analysis of the ATP-concentration-dependence indicates the presence of a rate-limiting monomolecular step that governs the process. From the temperature-dependence of this process, studied at saturating concentrations of ATP, an activation energy of 44 kJ/mol (10.6 kcal/mol) was measured. The spectral perturbation also occurs when cytochrome c oxidase is reconstituted into artificial phospholipid vesicles, with equilibria and kinetics similar to those observed with the soluble enzyme. Mixing ATP with soluble oxidized cyanide-bound cytochrome c oxidase induces a different spectral perturbation, and the apparent affinity of ATP for the enzyme is substantially increased. There is no absolute specificity for ATP, because EGTA, inositol hexakisphosphate, sulphate and phosphate are all able to induce an identical spectral perturbation with the same kinetics, although the value of the apparent Kd is different for the various anions. The presence of Mg2+ ions decreases, in a saturation-dependent fashion, the apparent affinity of cytochrome c oxidase for ATP. The absorbance change can be correlated to the displacement of the Ca2+ bound to cytochrome c oxidase.  相似文献   

18.
Asymmetry of tyrosyl-tRNA synthetase in solution   总被引:2,自引:0,他引:2  
W H Ward  A R Fersht 《Biochemistry》1988,27(3):1041-1049
The tyrosyl-tRNA synthetase from Bacillus stearothermophilus crystallizes as a symmetrical dimer with each subunit having a complete active site. The enzyme-substrate complexes, however, are known to be asymmetrical in solution because the enzyme exhibits half-of-the-sites activity by binding tightly only 1 mol of tyrosine or 1 mol of tyrosyl adenylate per mole of dimer. Evidence is now presented that the unligated enzyme is also asymmetrical in solution. Symmetry was investigated by construction of heterodimers containing one full-length subunit and one truncated subunit, allowing the introduction of different mutations into each monomer. Each dimer is active at only one site, but the site used is randomly distributed between the subunits. Each heterodimer thus consists of two equal populations, one activating tyrosine at a full-length subunit and the other at the truncated subunit. No detectable interconversion is found between active and inactive sites over several minutes either in the absence of substrates or when the enzyme is turning over in the steady state. Kinetic evidence implies that wild-type enzyme is inherently asymmetrical even in the absence of substrate.  相似文献   

19.
An Acinetobacter sp., strain CNU961, with a higher tolerance to phenol was isolated, and identified through a set of taxonomic studies and a genetic complementation test. Enzymatic and mutagenic studies found that the strain dissimilate phenol by hydroxylation to catechol followed by an ortho-ring cleavage pathway to further mineralize it. The phenol hydroxylase, which is an inducible enzyme and requires NADPH for optimum activity, was not inhibited by phenol at concentrations up to 0.5 mM. The different kinetic behaviors of the enzyme activities on NADPH and on phenol reflected that the phenol hydroxylase of strain CNU961 is a multisubunit allosteric enzyme consisting of heterogeneous polypeptides.  相似文献   

20.
Arginyl residues in the NADPH-binding sites of phenol hydroxylase   总被引:1,自引:0,他引:1  
Phenol hydroxylase was inactivated by the arginine reagents 2,3-butanedione, 1,2-cyclohexanedione, and phenylglyoxal. The cosubstrate NADPH, as well as NADP+ and several analogues thereof, protected the enzyme against inactivation. Phenol did not protect the activity against any of the reagents used, nor did modification by 2,3-butanedione affect the binding of phenol. We propose the presence of arginyl residues in the binding sites for the adenosine phosphate part of NADPH.  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