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 62 毫秒
1.
Horse-liver alcohol dehydrogenase has been dissociated and denatured by 6 M guanidinium hydrochloride. Removal of the denaturant under optimum conditions of the solvent leads to partial reactivation. The concentrations of the enzyme, as well as the coenzyme (NAD+), and Zn2+, affect the reactivation significantly, since high concentrations promote the formation of inactive aggregation products. Analyzing the kinetics of reactivation and reassociation, conditions far from equilibrium of dissociation-association provide maximum yields (approximately 70%). The sigmoidal kinetic traces suggest a superposition of first-order transconformation and second-order association reactions; the latter are corroborated by the concentration dependence of the reactivation reaction. The coenzyme, NAD+, has no influence on the kinetics of reactivation. Addition of Zn2+ leads to a significant decrease of the rate and yield of reactivation. The process of renaturation, as reflected by the regain of native fluorescence shows complex kinetics: rapid relaxations are followed by slower first-order and second-order processes which parallel reactivation.  相似文献   

2.
Enzyme protein fluorescence of di-furylacryloyl-glyceraldehyde-3-phosphate dehydrogenase (di-FA-GPDH:lambda max.excitation 290 nm, lambda max.emission 338 nm) is quenched about 28% on saturation with NAD+. Results of fluorometric titration of di-FA-GPDH with NAD+ suggest the presence of two tight and two loose coenzyme binding sites (Kdiss. 0.1 and 6.0 microM, respectively). Initial rates of the NAD(+)-dependent reaction of di-FA-GPDH with arsenate and phosphate and of mono-FA-GPDH with phosphate have been determined at varying coenzyme concentrations. The data suggest that binding of NAD+ at the tight sites does not activate the acyl group for its reaction with the acceptor (phosphate or arsenate). The group transfer reaction is dependent only on NAD+ binding to the loose sites, which carry the acyl group. The large difference in the NAD+ binding affinity to the two types of sites and their different effects on the group transfer reaction impart a sigmoidal shape to the rate versus [NAD+] plots. The sigmoidicity is abolished if the reactive SH groups at the unacylated sites are blocked by carboxymethylation.  相似文献   

3.
Thermal unfolding parameters were determined for a two-domain tetrameric enzyme, phosphorylating D-glyceraldehyde-3-phosphate dehydrogenase (GAPDH), and for its isolated NAD(+)-binding domain. At pH 8.0, the transition temperatures (t(max)) for the apoforms of the native Bacillus stearothermophilus GAPDH and the isolated domain were 78.3 degrees C and 61.9 degrees C, with calorimetric enthalpies (DeltaH(cal)) of 4415 and 437 kJ/mol (or 30.7 and 22.1 J/g), respectively. In the presence of nearly saturating NAD(+) concentrations, the t(max) and the DeltaH(cal) increased by 13.6 degrees C and by 2365 kJ/mol, respectively, for the native apoenzyme, and by 2.8 degrees C and 109 kJ/mol for the isolated domain. These results indicate that interdomain interactions are essential for NAD(+) to produce its stabilizing effect on the structure of the native enzyme. The thermal stability of the isolated NAD(+)-binding domain increased considerably upon transition from pH 6.0 to 8.0. By contrast, native GAPDH exhibited greater stability at pH 6.0; similar pH-dependencies of thermal stability were displayed by GAPDHs isolated from rabbit muscle and Escherichia coli. The binding of NAD(+) to rabbit muscle apoenzyme increased t(max) and DeltaH(cal) and diminished the widths of the DSC curves; the effect was found to grow progressively with increasing coenzyme concentrations. Alkylation of the essential Cys149 with iodoacetamide destabilized the apoenzyme and altered the effect of NAD(+). Replacement of Cys149 by Ser or by Ala in the B. stearothermophilus GAPDH produced some stabilization, the effect of added NAD(+) being basically similar to that observed with the wild-type enzyme. These data indicate that neither the ion pairing between Cys149 and His176 nor the charge transfer interaction between Cys149 and NAD(+) make any significant contribution to the stabilization of the enzyme's native tertiary structure and the accomplishment of NAD(+)-induced conformational changes. The H176N mutant exhibited dramatically lower heat stability, as reflected in the values of both DeltaH(cal) and t(max). Interestingly, NAD(+) binding resulted in much wider heat capacity curves, suggesting diminished cooperativity of the unfolding transition.  相似文献   

4.
Both cytosolic and mitochondrial aspartate transaminase can be resolved of pyridoxal phosphate. The resulting apoenzymes still bind individual structural components of the coenzyme. The separate contributions of coenzyme components to protein thermal stability have been independently assessed for phosphate ions (Pi) and for the pyridoxal or pyridoxamine components of the coenzyme. 31P NMR and differential scanning calorimetry reveal that the thermodynamic contributions of binding are not additive and are dissimilar for the two isozymes. High and low affinity sites for Pi binding are present in both apoenzymes with only the low affinity site being present in the holoenzyme forms. The contribution of both bound phosphates to increasing temperatures (Tm) and enthalpies (delta Hd) of denaturation differ between the isozymes and within sites. In either isozyme occupancy of the high affinity site by Pi produces only a 4- or 5- degree increase in the Tm value with respect to Pi-free apoenzyme. By contrast, in the mitochondrial apoenzyme, the presence of Pi at the second low affinity site increases the calorimetric parameters from Tm = 47 degrees C and delta Hd = 4.7 cal g-1 to Tm = 62 degrees C and delta Hd = 7 cal g-1. For cytosolic apoenzyme the respective changes are from 66 to 69.5 degrees C and 5.2 to 5.8 cal g-1. Addition of pyridoxal, but not pyridoxamine, displaces the high affinity Pi in both apoenzymes. This shows that the pyridine ring and Pi groups of pyridoxal-P bind exclusive of each other when they are not covalently linked as an ester, as in the coenzyme. The observation has been exploited as a method to prepare completely dephosphorylated mitochondrial apoenzyme. Electrostatic effects, structural differences in the phosphate binding pockets, and steric effects can be invoked to account for the Pi and pyridine binding behavior in the two proteins.  相似文献   

5.
The actions of Dexon on the NADH-ferricyanide oxidoreductase and the NADPH oxidase system of electron transfer particles (ETP) from beef heart as well as on the NADPH-cytochrome c oxidoreductase from brewer's yeast (Saccharomyces carlsbergensis Hansen) were investigated. The inhibition of the NADH dehydrogenase activity of ETP and that of the yeast enzyme correspond with respect to the following characteristics: 1) increase in the inhibition, 2) enhancement of the Dexon sensitivity by one order of magnitude after preincubation in the presence of NAD(P)H, 3) irreversibility of the inhibition, 4) no detectable changes in the spectral properties and in coenzyme activity of FMN after acid extraction from Dexon-treated enzyme. The inhibition of the NADH dehydrogenase activity of ETP is diminished by both NAD+ and FMN. However, no interaction of Dexon with NAD(P)H or FMN could be detected in the absence of enzyme or apoenzyme. The concentration of half-inhibition by Dexon for the yeast enzyme corresponds with its FMN concentration. It is proposed that both apoenzyme, NAD(P)H and FMN are involved in the interaction with Dexon. Possible mechanisms of binding are both complanar complexations of the ring systems and a triazene formation between FMNH2 and Dexon. The NADPH oxidase activity of the ETP is partly inhibited; the share inhibited by Dexon may represent the pathway via the transhydrogenase reaction.  相似文献   

6.
Rat liver S-adenosylhomocysteinase, a homotetramer, was resolved by treatment with acid ammonium sulfate into apoenzyme and NAD. The apoenzyme thus prepared retained a tetrameric structure but differed in the mobility on nondenaturing polyacrylamide gel electrophoresis. The inactive apoenzyme was reactivated upon incubation with NAD. The restoration of activity paralleled with the tight binding of NAD to apoenzyme, and full activity was obtained when 4 mol of NAD were bound per mol of apoenzyme. The kinetics of reconstitution were apparently biphasic and suggest the existence of two conformers in a slow equilibrium, one of which binds the coenzyme rapidly while the other does so very slowly, if at all. In addition to NAD, apoadenosylhomocysteinase tightly bound nicotinamide hypoxanthine dinucleotide, 3-acetylpyridine adenine dinucleotide and nicotinic acid-adenine dinucleotide. NADP was not bound. Catalytic activity was found only with the enzyme reconstituted with NAD or nicotinamide hypoxanthine dinucleotide. The spectral change observed on interaction of apoadenosylhomocysteinase with NAD was similar to those seen with adenine nucleotides, and was largely approximated by the addition of dioxane to aqueous solutions of adenine nucleotides. By comparison of the difference spectra, it is suggested that the adenine portion of the coenzyme is bound in the hydrophobic pocket of the protein, and that the binding is accompanied by perturbation of tryptophan residue of the protein.  相似文献   

7.
D Eby  M E Kirtly 《Biochemistry》1976,15(10):2168-2171
Using NAD analogues as ligands, the structural requirements for negative cooperativity in binding to rabbit muscle glyceraldehyde-3-phosphate dehydrogenase were examined. Although the affinity of nicotinamide hypoxanthine dinucleotide is considerably lower than that of NAD+, it also binds to the enzyme with negative cooperatively. Two pairs of nicotinamide hypoxanthine dinucleotide binding sitess were distinguished, one pair having an affinity for the analogue which is 15 times that of the second pair. Negative cooperativity is also found in the Km values for the analogue. Thus modification of the adenine ring of NAD+ to hypoxanthine does not abolish negative cooperativity in coenzyme binding. Adenosine diphosphoribose binding to the same enzyme shows neither positive nor negative cooperativity, indicating that cooperativity apparently requires an intact nicotinamide ring in the coenzyme structure, under the conditions of these experiments. Occupancy of the nicotinamide subsite of the coenzyme binding site is not necessary for half-of-sites reactivity of alkylating or acylating compounds (Levitzki, A. (1974), J. Mol, Biol. 90, 451-458). However, it can be important in the negative cooperativity in ligand binding, as illustrated by adenosine diphosphoribose which fails to exhibit negative cooperativity. Occupancy of the adenine subsite by adenine is important for stabilization of the enzyme against thermal denaturation. Whether the stabilization is due to an altered conformation of the subunits or stabilization of the preexisting structure of the apoenzyme cannot be determined from these studies. However, nicotinamide hypoxanthine dinucleotide does not contribute to enzyme stability although it serves as a substrate and shows negative cooperativity.  相似文献   

8.
Crystals of apo- and holo-D-glyceraldehyde-3-phosphate dehydrogenase from the tail muscle of the Mediterranean lobster Palinurus vulgaris, previously found to be suitable for single crystal microspectrophotometric studies of catalytic activity in the crystalline state, have been examined by x-ray crystallography. The two forms are isomorphous, space group C 2 with cell dimensions a=128.4 A, b=99.9 A, c=80.3 A,beta=113.4 degrees. These data are consistent with a molecular weight of 73,000 in the crystallographic asymmetric unit, indicating that the tetrameric molecule possesses an exact 2-fold axis both in the presence and in the absence of NAD+. Analysis of the intensity distribution of conventional x-ray precession photographs shows that two further noncrystallographic diads are present and that the molecule has the 2 pseudo 22 symmetry found in other D-glyceraldehyde-3-phosphate dehydrogenases. Binding of NAD+ to apoenzyme in solution, at 25 degrees C, is anticooperative and it can be satisfactorily described by assuming two classes of coenzyme binding sites.  相似文献   

9.
1. Glyceraldehyde-3-phosphate dehydrogenase from bacillus stearothermophilus can be extensively succinylated in the presence of substrates and coenzyme without appreciable loss of activity. 2. The apoenzyme in the absence of substrates is rapidly inhibited by small amounts of succinic anhydride. NAD+, glyceraldehyde-3-phosphate and inorganic phosphate all afford protection from inhibition, and inhibition is slowly reversed in the presence of pyrophosphate at pH 8.5. 3. Kinetic and spectral studies have shown that the specific inhibition is associated with the succinylation of the aliphatic hydroxyl group of a serine or threonine residue. 4. The residue specifically succinylated has been identified as one of the two threonine residues, most probably Thr-150, adjacent to the activ-site cysteine residue in the primary structure. Its unusual reactivity is discussed in relation to the three-dimensional structure of the enzyme. 5. A second residue, a lysine homologous with Lys-212 in the pig muscle enzyme, can be succinylated in both holoenzyme and apoenzyme with no detectable effect upon the enzymic activity.  相似文献   

10.
D-beta-Hydroxybutyrate dehydrogenase D-3-hydroxybutyrate: NAD+ oxidoreductase, EC 1.1.1.30), a phosphatidylcholine-requiring enzyme, was irreversibly inactivated by a water-soluble carbodiimide, 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDAC) or a hydrophobic carbodiimide, N,N'-dicyclohexylcarbodiimide (DCCD). The inactivation is pseudo-first-order with a kinetic stoichiometry of about 1. Phospholipid-free apoenzyme was more sensitive towards these reagents than reconstituted phospholipid-enzyme or membrane-bound enzyme forms. Reduced coenzyme (NADH) protected the enzyme against the inactivation, while oxidized coenzyme (NAD+) in presence of 2-methylmalonate (a pseudo-substrate) gave a better protection. It was found that the phospholipid-free apoenzyme bound about 1 mol [14C]DCCD. This incorporation was prevented by EDAC, indicating that both reagents react at the same site. [14C]Glycine ethyl ester, a nucleophilic compound which reacts specifically with the carboxylcarbodiimide derivative was incorporated to the enzyme (1 mol [14C]glycine ethyl ester per polypeptide chain), whatever its form, in the presence of DCCD or EDAC. These results indicate the presence of one carboxyl group probably located at or near the coenzyme-binding site and near the interacting domain of the enzyme with phospholipid.  相似文献   

11.
The formation of binary complexes between sturgeon apoglyceralddhyde-3-phosphate dehydrogenase, coenzymes (NAD+ and NADH) and substrates (phosphate, glyceraldehyde 3-phosphate and 1,3-bisphosphoglycerate) has been studied spectrophotometrically and spectrofluorometrica-ly. Coenzyme binding to the apoenzyme can be characterized by several distinct spectroscopic properties: (a) the low intensity absorption band centered at 360 nm which is specific of NAD+ binding (Racker band); (b) the quenching of the enzyme fluorescence upon coenzyme binding; (c) the quenching of the fluorescence of the dihydronicotinamide moiety of the reduced coenzyme (NADH); (D) the hypochromicity and the red shift of the absorption band of NADH centered at 338 nm; (e) the coenzyme-induced difference spectra in the enzyme absorbance region. The analysis of these spectroscopic properties shows that up to four molecules of coenzyme are bound per molecule of enzyme tetramer. In every case, each successively bound coenzyme molecule contributes identically to the total observed change. Two classes of binding sites are apparent at lower temperatures for NAD+ Binding. Similarly, the binding of NADH seems to involve two distinct classes of binding sites. The excitation fluorescence spectra of NADH in the binary complex shows a component centered at 260 nm as in aqueous solution. This is consistent with a "folded" conformation of the reduced coenzyme in the binary complex, contradictory to crystallographic results. Possible reasons for this discrepancy are discussed. Binding of phosphorylated substrates and orthophosphate induce similar difference spectra in the enzyme absorbance region. No anticooperativity is detectable in the binding of glyceraldehyde 3-phosphate. These results are discussed in light of recent crystallographic studies on glyceraldehyde-3-phosphate dehydrogenases.  相似文献   

12.
R Rudolph  I Heider  R Jaenicke 《Biochemistry》1977,16(25):5527-5531
Dissociation and deactivation of the H4 and M4 isoenzymes of lactic dehydrogenase in strong denaturants may be reversed with a yield of reactivation up to 100%. The products of reconstitution are indistinguishable from the native enzymes as far as the Michaelis constants and the dissociation constants for substrate and coenzyme as well as spectral and hydrodynamic properties are concerned. The presence of NAD+ and NADH does not affect either the conformational state of the product of reconstitution, or the kinetics of reactivation, using the pure apoenzymes as a reference. At 20 degrees C the kinetics of reactivation for LDH-M4 in the presence and absence of coenzyme may be quantitatively described by a second-order rate equation (k2 = 23.4 +/- 2.6 mM-1S-1) while LDH-H4 is characterized by a uni-bimolecular reaction sequence (k1 = 1.45 +/- 0.45 X 10(-3)-S-1, k2 = 5 +/- 1 mM-1S-1), in agreement with earlier observations (Rudolph, R., et al. (1977), Biochemistry 16, 3384-3390). Regarding the influence of temperature on the rate of reactivation no significant anomalies are detectable within the range of 0-25 degrees C. The (apparent) activation energies, taken from the linear Arrhenius plots, are 58 kcal/mol for the association reaction of LDH-M4, and 41 kcal/mol for the transconformation reaction of LDH-H4.  相似文献   

13.
UDPglucose 4-epimerase from Kluyveromyces fragilis was completely inactivated by diethylpyrocarbonate following pseudo-first order reaction kinetics. The pH profile of diethylpyrocarbonate inhibition and reversal of inhibition by hydroxylamine suggested specific modification of histidyl residues. Statistical analysis of the residual enzyme activity and the extent of modification indicated modification of 1 essential histidine residue to be responsible for loss in catalytic activity of yeast epimerase. No major structural change in the quarternary structure was observed in the modified enzyme as shown by the identical elution pattern on a calibrated Sephacryl 200 column and association of coenzyme NAD to the apoenzyme. Failure of the substrates to afford any protection against diethylpyrocarbonate inactivation indicated the absence of the essential histidyl residue at the substrate binding region of the active site. Unlike the case of native enzyme, sodium borohydride failed to reduce the pyridine moiety of the coenzyme in the diethylpyrocarbonate-modified enzyme. This indicated the presence of the essential histidyl residue in close proximity to the coenzyme binding region of the active site. The abolition of energy transfer phenomenon between the tryptophan and coenzyme fluorophore on complete inactivation by diethylpyrocarbonate without any loss of protein or coenzyme fluorescence are also added evidences in this direction.  相似文献   

14.
Initial-rate studies were made of the oxidation of L-glutamate by NAD+ and NADP+ catalysed by highly purified preparations of dogfish liver glutamate dehydrogenase. With NAD+ as coenzyme the kinetics show the same features of coenzyme activation as seen with the bovine liver enzyme [Engel & Dalziel (1969) Biochem. J. 115, 621--631]. With NADP+ as coenzyme, initial rates are much slower than with NAD+, and Lineweaver--Burk plots are linear over extended ranges of substrate and coenzyme concentration. Stopped-flow studies with NADP+ as coenzyme give no evidence for the accumulation of significant concentrations of NADPH-containing complexes with the enzyme in the steady state. Protection studies against inactivation by pyridoxal 5'-phosphate indicate that NAD+ and NADP+ give the same degree of protection in the presence of sodium glutarate. The results are used to deduce information about the mechanism of glutamate oxidation by the enzyme. Initial-rate studies of the reductive amination of 2-oxoglutarate by NADH and NADPH catalysed by dogfish liver glutamate dehydrogenase showed that the kinetic features of the reaction are very similar with both coenzymes, but reactions with NADH are much faster. The data show that a number of possible mechanisms for the reaction may be discarded, including the compulsory mechanism (previously proposed for the enzyme) in which the sequence of binding is NAD(P)H, NH4+ and 2-oxoglutarate. The kinetic data suggest either a rapid-equilibrium random mechanism or the compulsory mechanism with the binding sequence NH4+, NAD(P)H, 2-oxoglutarate. However, binding studies and protection studies indicate that coenzyme and 2-oxoglutarate do bind to the free enzyme.  相似文献   

15.
Ammonium sulfate, a typical component of crystallization media of proteins, stabilizes an inactive conformation of pig muscle glyceraldehyde-3-phosphate dehydrogenase. In fact, in the presence of ammonium sulfate the reconstitution of the catalytically active holoenzyme from the apoenzyme and NAD is not instantaneous, as in the case of enzymes from Bacillus stearothermophilus and the Mediterranean lobster Palinurus vulgaris. With pig muscle enzyme, at pH 6.0, the time course of formation of the characteristic Racker band can be monitored by a rapid mixing stopped flow technique. Activation follows a single exponential curve with a rate constant independent of the concentration of both NAD and protein and, therefore, appears to be limited by a slow protein isomerization (k = 7 +/- 2 s-1). Accordingly, when the apoenzyme is simultaneously exposed to NAD and either glyceraldehyde 3-phosphate or 1,3-bisphosphoglycerate, the ensuing reactions (the redox and the acylation steps, respectively) are kinetically limited by the same protein isomerization. At pH 7.0 and 8.0, however, two among the four active sites react with NAD at an unmeasurably high rate, while the other two sites behave as they do at pH 6.0. When the pig muscle apoenzyme is preincubated and allowed to react with either glyceraldehyde 3-phosphate or 1,3-bisphosphoglycerate before the rapid mixing with NAD, both the redox reaction and the NAD-dependent activation of apo-acyl-enzyme toward arsenolysis become unmeasurably fast. Similarly, when the sulfate in the medium is replaced by ions such as phosphate and citrate, the reconstitution of the active holoenzyme is practically instantaneous. Thus, the slow protein isomerization observed in the presence of sulfate and abolished by competing substrates and anions is diagnostic of a structural state of the pig muscle apoenzyme, which is induced by sulfate ions bound within the enzyme active site.  相似文献   

16.
Inactivation of apo-glyceraldehyde-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate: NAD+ oxidoreductase(phosphorylating) (EC 1.2.1.12) from rat skeletal muscle at 4 degrees C in 0.15 M NaC1, 5 mM EDTA, 4 mM 2-mercaptoethanol pH 7.2 is a first-order reaction. The rate constant of inactivation depends on protein concentration. With one molecule of NAD bound per tetrameric enzyme, a 50 per cent loss in activity is observed and the rate constant of inactivation becomes independent of the protein concentration over a 30-fold range. Two moles of NAD bound per mole of enzyme fully protect it against inactivation. NADH affords a cooperative effect on enzyme structure similar to that of NAD. Inactivation of 7.8 S apoenzyme is reflected in its dissociation into 4.8-S dimers. In the case of enzyme-NAD1 complex, no direct relationship between the extent of inactivation and dissociation is observed, suggesting that these two processes do not occur simultaneously; we may say that dissociation is slower than inactivation. A mechanism in which the rate-limiting step for inactivation is a conformational change in the tetramer occurring prior to dissociation and affecting only the structure of the non-liganded dimer, is consistent with the experimental observations. Inorganic phosphate protects apoenzyme against inactivation. Its effect is shown to be due to the anion binding at specific sites on the protein with a dissociation constant of 2.6 plus or minus 0.4 mM. The NaC1-induced cold inactivation of glyceraldehyde-phosphate dehydrogenase is fully reversible at 25 degrees C in the presence of 20 mM dithiothreitol and 50 mM inorganic phosphate. The rate of reactivation is independent of protein concentration. Inactivated enzyme retains the ability to bind specific antibodies produced in rabbits, but diminishes its precipitating capability.  相似文献   

17.
Reactivation of substrate-inactivated brain glutamate decarboxylase   总被引:8,自引:0,他引:8  
The effects of ATP and inorganic phosphate (Pi) on the reactivation of glutamate apodecarboxylase by its cofactor pyridoxal-5'-phosphate (pyridoxal-P) was studied. Apoenzyme was prepared by preincubation with glutamate. Apoenzyme prepared with glutamate alone was reactivated slowly and incompletely by adding a saturating concentration of pyridoxal-P (20 microM). Reactivation was slightly enhanced by 1-10 mM Pi. Reactivation by pyridoxal-P plus Pi was greatly enhanced by the presence of low concentrations (less than 100 microM) of ATP during the preparation of apoenzyme with glutamate. Reactivation was much lower if Pi was omitted. Enhancement of reactivation by ATP was due to its effect during apoenzyme formation, since ATP did not enhance reactivation if added only during reactivation and since the enhancing effect persisted after the removal of free ATP by chromatography on Sephadex G-25 after apoenzyme preparation and before reactivation. Reactivation was inhibited by high concentrations of ATP (greater than 100 microM), possibly by competition of ATP for the cofactor binding site. Four factors (glutamate, pyridoxal-P, ATP, and Pi) control a cycle of inactivation and reactivation that appears to be important in the regulation of brain glutamate decarboxylase.  相似文献   

18.
K Mori  T Toraya 《Biochemistry》1999,38(40):13170-13178
The mechanism of reactivation of diol dehydratase by its reactivating factor was investigated in vitro by using enzyme. cyanocobalamin complex as a model for inactivated holoenzyme. The factor mediated the exchange of the enzyme-bound, adenine-lacking cobalamins for free, adenine-containing cobalamins through intermediate formation of apoenzyme. The factor showed extremely low but distinct ATP-hydrolyzing activity. It formed a tight complex with apoenzyme in the presence of ADP but not at all in the presence of ATP. Incubation of the enzyme.cyanocobalamin complex with the reactivating factor in the presence of ADP brought about release of the enzyme-bound cobalamin, leaving the tight apoenzyme-reactivating factor complex. Although the resulting complex was inactive even in the presence of added adenosylcobalamin, it dissociated by incubation with ATP, forming the apoenzyme, which was reconstitutable into active holoenzyme with added coenzyme. Thus, it was established that the reactivation of the inactivated holoenzyme by the factor in the presence of ATP and Mg2+ takes place in two steps: ADP-dependent cobalamin release and ATP-dependent dissociation of the apoenzyme.factor complex. ATP plays dual roles as a precursor of ADP in the first step and as an effector to change the factor into the low-affinity form for diol dehydratase. The enzyme-bound adenosylcobalamin was also susceptible to exchange with free adeninylpentylcobalamin, although to a much lesser degree. The mechanism for discrimination of adenine-containing cobalamins from adenine-lacking cobalamins was explained in terms of formation equilibrium constants of the cobalamin.enzyme.reactivating factor ternary complexes. We propose that the reactivating factor is a new type of molecular chaperone that participates in reactivation of the inactivated enzymes.  相似文献   

19.
On- and off-velocity constants for NADH and NAD+ binding to liver alcohol dehydrogenase in the pH range 10-12 have been determined by stopped-flow kinetic methods. The results are consistent with previously reported equilibrium binding data and proposals attributing the main effects of pH on coenzyme binding to ionization of Lys-228 and zinc-bound water. Deprotonation of the group identified as Lys-228 decreases the NADH and NAD+ association rates by a factor exceeding 20 and has no detectable effect on the coenzyme dissociation rates in the examined pH range. Ionization of the group identified as zinc-bound water causes a 3-fold increase of the rate of NADH dissociation from the enzyme, and decreases the rate of NAD+ dissociation by a factor of 200. The NADH and NAD+ association rates are decreased by a factor of 30 and 5, respectively. The observed effects of pH can be rationalized in terms of electrostatic interactions of the ionizing groups with the charges present on the coenzyme molecules and lend support to the idea that binding of the coenzyme nicotinamide ring occurs subsequent to binding of the AMP portion of the coenzyme.  相似文献   

20.
Inactivation of apo-glyceraldehyde-3-phosphate dehydrogenase from rat skeletal muscle in the presence of butanedione is the result of modification of one arginyl residue per subunit of the tetrameric enzyme molecule. The loss of activity follows pseudo-first-order kinetics. NAD+ increases the apparent first-order rate constant of inactivation. The effect of NAD+ on the enzyme inactivation is cooperative (Hill coefficient = 2.3--3.2). Glyceraldehyde 3-phosphate protected the holoenzyme against inactivation, decreasing the rate constant of the reaction. At saturating concentrations of substrate the protection was complete. The Hill plot demonstrates that the effect is cooperative. This suggests that subunit interactions in the tetrameric holoenzyme molecule may affect the reactivity of the essential arginyl residues. In contrast, glyceraldehyde 3-phosphate had no effect on the rate of inactivation of the apoenzyme in the presence of butanedione. 100 mM inorganic phosphate protected both the apoenzyme and holoenzyme against inactivation. The involvement of the microenvironment of the arginyl residues in the functionally important conformational changes of the enzyme is discussed.  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