首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The yeast plasma membrane proton-pumping ATPase forms a phosphorylated intermediate during the hydrolysis of ATP. The fraction of enzyme phosphorylated during steady-state ATP hydrolysis was studied as a function of substrate concentration (MgATP), Mg2+ concentration, and pH. The dependence of the fraction of enzyme phosphorylated on the concentration of MgATP is sigmoidal, and the isotherms can be fit with parameters and mechanisms similar to those used to describe ATP hydrolysis. The isotherm is significantly more sigmoidal at pH 5.5 than at pH 6.0, with the limiting percentage (100.mol of phosphate/mol of enzyme) of enzyme phosphorylated being 70% and 6%, respectively, at the two pH values. The maxima in the steady-state rate of ATP hydrolysis occur at higher concentrations of Mg2+ and higher pH than the maxima in the fraction of enzyme phosphorylated. This suggests that the rate-determining step for ATP hydrolysis is different from that for enzyme phosphorylation and the hydrolysis of phosphoenzyme is enhanced by Mg2+ and high pH. The rate of phosphoenzyme formation was investigated with the quenched-flow method, but only a lower bound of 140 s-1 could be obtained for the rate constant at MgATP concentrations greater than 2.5 mM. Since the turnover number for ATP hydrolysis under similar conditions is 14 s-1, the rate-determining step in ATP hydrolysis occurs after enzyme phosphorylation.  相似文献   

2.
The time course of ATP-Pi exchange which is catalyzed by the isolated chloroplast ATP synthase in phospholipid vesicles was studied. The following observations were made. (i) The onset of 32Pi incorporation into ATP lags behind ATP hydrolysis. The lag lasts for about 2 min at 37 degrees C and is followed by a steady-state rate which is constant for more than 30 min. Under the same experimental conditions, ATP hydrolysis shows an initial burst followed by a constant, slower rate. (ii) The initial lag is independent of Mg-ATP concentration in the range 0.2-5 mM and of the presence of ADP. In contrast, the steady-state rate of ATP-Pi exchange has an apparent Km of 0.3 mM for Mg-ATP and is stimulated by ADP. (iii) Increasing the temperature from 30 to 45 degrees C decreases the lag from 6 min to zero. The steady-state rate of ATP-Pi exchange is affected to a much smaller extent by the temperature in this range. (iv) The lag is insensitive to valinomycin or tetraphenylboron, while the steady-state rate is partially inhibited. Nigericin and protonophores affect both the lag and steady-state rate. (v) ATP-induced membrane potential formation, as followed by oxonol VI, does not correlate with the lag in its kinetics and temperature dependence. ATP-induced pH gradient formation could not be detected in the proteoliposome system. (vi) Light-triggered ATP-Pi exchange in chloroplasts shows essentially the same time course as the proteoliposome system, but the lag lasts for only about 20 s at room temperature and is unaffected by a preexisting proton gradient. These results suggest that the initial lag in ATP-Pi exchange does not reflect the time required for the buildup of a protomotive force (delta - mu H+) nor the time required to produce ADP. It is suggested, therefore, that the lag reflects an internal autocatalytic conformational change in the ATP-synthase complex which is initiated by ATP hydrolysis and which converts the enzyme from an "exclusive ATPase state" to a "reversible ATP-synthase state". This slow transition is not directly coupled to a trans-membrane pH or potential gradient.  相似文献   

3.
Pre-steady-state studies of the isocitrate dehydrogenase reaction show that the rate constant for the hydride-transfer step is above 990s-1, and that both subunits of the enzyme are simulataneously active. After the fast formation of NADPH in amounts equivalent to the enzyme subunit concentration, the rate of NADPH formation is equal to the steady-state rate if the enzyme has been preincubated with isocitrate and Mg2+. If the enzyme has been preincubated with NADP+ and Mg2+, in 0.05 M-triethanolamine chloride buffer, pH 7.0, with the addition of 0.1 M-NaCl, the amount of NADPH formed in the fast phase is only 60% of the enzyme subunit concentration, and the turnover rate is at first lower than the steady-state rate. In 0.05 M-triethanolamine chloride buffer, pH 7.0, if the enzyme is preincubated with NADP+ or NADPH, the turnover rate increases 3-fold to reach the steady-state rate after about 5 s. Preincubation of the enzyme with isocitrate and Mg2+ abolishes this lag phase, the steady-state rate being reached at once. It is suggested that the enzyme exists in at least two conformational forms with different activities, and that the lag phase represents the transition (k = 0.4s-1) from a form with low activity to the fully active enzyme, induced by the binding of isocitrate and Mg2+.  相似文献   

4.
Steady-state kinetics for the hydrolysis of benzoylcholine (BzCh) and benzoylthiocholine (BzSCh) by wild-type human butyrylcholinesterase (BuChE) and by the peripheral anionic site mutant D70G were compared. kcat/Km for the hydrolysis of BzSCh was 17-fold and 32-fold lower than that for hydrolysis of BzCh by wild-type and D70G, respectively. The rate-limiting step for hydrolysis of BzCh was deacylation, whereas acylation was rate-limiting for hydrolysis of BzSCh. Wild-type enzyme and the D70G mutant were found to reach steady-state velocity slowly with BzCh as the substrate. At pH 6, the approach to steady-state for both enzymes consisted of a mono-exponential acceleration upon which a set of damped oscillations was superimposed. From pH 7 to 8.5, the approach to steady-state consisted of a simple exponential acceleration. The damped oscillations were analyzed by both a numerical approximation and simulation based on a theoretical model. BuChE-catalyzed hydrolysis of the thiocholine analogue of BzCh showed neither lags nor oscillations, under the same conditions. The frequency and amplitude of the damped oscillations decreased as the BzCh concentration increased. The apparent induction time for the exponential portion of the lag was calculated from the envelope of the damped oscillations or from the smooth lag. Wild-type BuChE showed a hyperbolic increase in induction time as the BzCh concentration increased (tau max = 210 s at pH 6.0). However, the induction time for D70G was constant over the whole range of BzCh concentrations (tau max = 60 s at pH 6.0). Thus, the induction time does not conform to a simple hysteretic model in which there is a slow conformational transition of the enzyme from an inactive form E to an active form E'. No pH-dependence of the induction time was found between pH 6.0 and 8.5 in sodium phosphate buffers of various concentrations (from 1 mm to 1 m). However, increasing the pH tended to abolish the oscillations (increase the damping factor). This effect was more pronounced for D70G than for wild-type. Although the lyotropic properties of phosphate change from chaotropic at pH 6.0 to kosmotropic at pH > 8.0, no effect of phosphate concentration on the oscillations was noticed at the different pH values, suggesting that the oscillations are not related to a pH-dependent Hofmeister effect of phosphate ions. Simulation and theoretical analysis of the oscillatory behaviour of the approach to the steady-state for BuChE led us to propose a model for the hysteresis of BuChE with BzCh. In this model, the substrate-free enzyme is present as an equilibrium mixture of two forms, E and E'. Substrate binds to E and E', but only Epsilon'S makes products. It is proposed that oscillations originate from a time-dependent change in the local concentration, solvation and/or conformation of substrate in the bulk solution. 1H-NMR measurements provided evidence for a slow equilibrium between two BzCh conformers. Binding of the conformationally preferred substrate conformer leads to products.  相似文献   

5.
UDP-galacturonic acid, the activated form of galacturonic acid (GalUA), is synthesized both de novo and by salvage pathways. The UDP-GalUA pyrophosphorylase gene involved in the salvage pathway has not been identified. Here we show that UDP-sugar pyrophosphorylase from Pisum sativum with a broad specificity has UDP-GalUA pyrophosphorylase activity. The enzyme catalyzed the formation of UDP-GalUA and pyrophosphate from GalUA 1-phosphate and UTP with an equilibrium constant value of 0.24. The recombinant UDP-sugar pyrophosphorylase had optimal pH of 6.0, and the apparent K(m) values for GalUA 1-phosphate, UTP, UDP-GalUA, and pyrophosphate were 2.27, 1.15, 0.70, and 1.26 mM, respectively. In the presence of inorganic pyrophosphatase, the recombinant enzyme produced UDP-GalUA in an 84% yield (based on the GalUA 1-phosphate substrate) on a preparative scale. Thus, this UDP-sugar pyrophosphorylase is useful for the highly efficient production of UDP-GalUA for studies on pectin biosynthesis.  相似文献   

6.
C Ghose  F M Raushel 《Biochemistry》1985,24(21):5894-5898
The reactions catalyzed by argininosuccinate synthetase have been examined by the use of static and dynamic quench techniques. The time course of the forward reaction (22 degrees C) at pH 8.0 is characterized by a "burst" of AMP formation upon quenching with acid that is equivalent to 0.59 mol of enzyme. The pre-steady-state rate is followed by a slower steady-state rate of 0.60 s-1. The rate constant for the transient phase is 9.7 s-1. The time course for the formation of argininosuccinate is linear and shows neither a "lag" nor a burst phase. These results have been interpreted to mean that the mechanism for the formation of argininosuccinate consists of at least two distinct chemical steps with the formation of citrulline adenylate as a reactive intermediate. In the presence of aspartate the rate constant for the formation of citrulline adenylate (6.2 s-1) from ATP and citrulline is 7 times faster than the rate of formation of argininosuccinate from aspartate and citrulline adenylate (0.9 s-1). This suggests that the second step is predominantly rate limiting. The rate constant for the formation of citrulline adenylate in the absence of enzyme-bound aspartate (0.01 s-1) is 600 times slower than when aspartate is present. This indicates that the binding of aspartate to the enzyme regulates the formation of the intermediate. These results are in complete accord with our previously published steady-state kinetic scheme showing sequential addition of substrates.  相似文献   

7.
The complete time course of the hydrolysis of p-nitrophenyl phosphate catalyzed by the low molecular weight (acid) phosphotyrosyl protein phosphatase from bovine heart was elucidated and analyzed in detail. Burst titration kinetics were demonstrated for the first time with this class of enzyme. At pH 7.0, 4.5 degrees C, a transient pre-steady-state "burst" of p-nitrophenol was formed with a rate constant of 48 s-1. The burst was effectively stoichiometric and corresponded to a single enzyme active site/molecule. The burst was followed by a slow steady-state turnover of the phosphoenzyme intermediate with a rate constant of 1.2 s-1. Product inhibition studies indicated an ordered uni-bi kinetic scheme for the hydrolysis. Partition experiments conducted for several substrates revealed a constant product ratio. Vmax was constant for these substrates, and the overall rate of hydrolysis was increased greatly in the presence of alcohol acceptors. An enzyme-catalyzed 18O exchange between inorganic phosphate and water was detected and occurred with kcat = 4.47 x 10(-3) s-1 at pH 5.0, 37 degrees C. These results were all consistent with the existence of a phosphoenzyme intermediate in the catalytic pathway and with the breakdown of the intermediate being the rate-limiting step. The true Michaelis binding constant Ks = 6.0 mM, the apparent Km = 0.38 mM, and the rate constants for phosphorylation (k2 = 540 s-1) and dephosphorylation (k3 = 36.5 s-1) were determined under steady-state conditions with p-nitrophenyl phosphate at pH 5.0 and 37 degrees C in the presence of phosphate acceptors. The energies of activation for the enzyme-catalyzed hydrolysis at pH 5.0 and 7.0 were 13.6 and 14.1 kcal/mol, respectively. The activation energy for the enzyme-catalyzed medium 18O exchange between phosphate and water was 20.2 kcal/mol. Using the available equilibrium and rate constants, an energetic diagram was constructed for the enzyme-catalyzed reaction.  相似文献   

8.
The steady-state kinetics of the oxidative decarboxylation of 6-phosphogluconate catalysed by 6-phosphogluconate dehydrogenase from sheep liver in triethanolamine and phosphate buffers (pH 7.0) have been reinvestigated. In triethanolamine buffer the enzyme is inhibited by high NADP+ concentrations in the presence of low fixed concentrations of 6-phosphogluconate. Data are consistent with an asymmetric sequential mechanism in which NADP+ and 6-phosphogluconate bind randomly and product release is ordered. The pathway through the enzyme--6-phosphogluconate complex appears to be preferred in triethanolamine buffer. Pre-steady-state studies of the oxidative decarboxylation reaction at pH 6.0-8.0 show that hydride transfer is greater than 900 s-1. After the fast formation of NADPH in amounts equivalent to about half of the enzyme-active-centre concentration, the rate of NADPH formation is equal to the steady-state rate. Two possible interpretations are considered. Rapid fluorescence measurements of the displacement of NADPH from its complex with the enzyme at pH 6.0 and 7.0 indicate that the dissociation of NADPH is fast (greater than 800 s-1) and cannot be the rate-limiting step in oxidative decarboxylation. Coenzyme binding studies at equilibrium have been extended to include the determination of the dissociation constants for the binary complexes of enzyme with NADPH and NADP+ at pH 6.0-8.0 and the dissociation constant for NADPH in the ternary enzyme--6-phosphogluconate--NADPH complex in triethanolamine buffer, pH 7.0.  相似文献   

9.
Sergienko EA  Jordan F 《Biochemistry》2002,41(12):3952-3967
Pyruvate decarboxylase from yeast (YPDC, EC 4.1.1.1) exhibits a marked lag phase in the progress curves of product (acetaldehyde) formation. The currently accepted kinetic model for YPDC predicts that, only upon binding of substrate in a regulatory site, a slow activation step converts inactive enzyme into the active form. This allosteric behavior gives rise to sigmoidal steady-state kinetics. The E477Q active site variant of YPDC exhibited hyperbolic initial rate curves at low pH, not consistent with the model. Progress curves of product formation by this variant were S-shaped, consistent with the presence of three interconverting conformations with distinct steady-state rates. Surprisingly, wild-type YPDC at pH < or =5.0 also possessed S-shaped progress curves, with the conformation corresponding to the middle steady state being the most active one. Reexamination of the activation by substrate of wild-type YPDC in the pH range of 4.5-6.5 revealed two characteristic transitions at all pH values. The values of steady-state rates are functions of both pH and substrate concentration, affecting whether the progress curve appears "normal" or S-shaped with an inflection point. The substrate dependence of the apparent rate constants suggested that the first transition corresponded to substrate binding in an active site and a subsequent step responsible for conversion to an asymmetric conformation. Consequently, the second enzyme state may report on "unregulated" enzyme, since the regulatory site does not participate in its generation. This enzyme state utilizes the alternating sites mechanism, resulting in the hyperbolic substrate dependence of initial rate. The second transition corresponds to binding a substrate molecule in the regulatory site and subsequent minor conformational adjustments. The third enzyme state corresponds to the allosterically regulated conformation, previously referred to as activated enzyme. The pH dependence of the Hill coefficient suggests a random binding of pyruvate in a regulatory and an active site of wild-type YPDC. Addition of pyruvamide or acetaldehyde to YPDC results in the appearance of additional conformations of the enzyme.  相似文献   

10.
1. The mechanisms of the reduction of oxaloacetate and of 3-fluoro-oxaloacetate by NADH catalysed by cytoplasmic pig heart malate dehydrogenase (MDH) were investigated. 2. One mol of dimeric enzyme produces 1.7+/-0.4 mol of enzyme-bound NADH when mixed with saturating NAD+ and L-malate at a rate much higher than the subsequent turnover at pH 7.5. 3. Transient measurements of protein and nucleotide fluorescence show that the steady-state complex in the forward direction is MDH-NADH and in the reverse direction MDH-NADH-oxaloacetate. 4. The rate of dissociation of MDH-NADH was measured and is the same as Vmax. in the forward direction at pH 7.5. Both NADH-binding sites are kinetically equivalent. The rate of dissociation varies with pH, as does the equilibrium binding constant for NADH. 5. 3-Fluoro-oxaloacetate is composed of three forms (F1, F2 and S) of which F1 and F2 are immediately substrates for the enzyme. The third form, S, is not a substrate, but when the F forms are used up form S slowly and non-enzymically equilibrates to yield the active substrate forms. S is 2,2-dihydroxy-3-fluorosuccinate. 6. The steady-state compound during the reduction of form F1 is an enzyme form that does not contain NADH, probably MDH-NAD+-fluoromalate. The steady-state compound for form F2 is an enzyme form containing NADH, probably MDH-NADH-fluoro-oxaloacetate. 7. The rate-limiting reaction in the reduction of form F2 shows a deuterium isotope rate ratio of 4 when NADH is replaced by its deuterium analogue, and the rate-limiting reaction is concluded to be hydride transfer. 8. A novel titration was used to show that dimeric cytoplasmic malate dehydrogenase contains two sites that can rapidly reduce the F1 form of 3-fluoro-oxaloacetate. The enzyme shows 'all-of-the-sites' behaviour. 9. Partial mechanisms are proposed to explain the enzyme-catalysed transformations of the natural and the fluoro substrates. These mechanisms are similar to the mechanism of pig heart lactate dehydrogenase and this, and the structural results of others, can be explained if the two enzymes are a product of divergent evolution.  相似文献   

11.
Regulation of Pyruvate Decarboxylase In Vitro and In Vivo   总被引:2,自引:0,他引:2  
Results presented in this paper strongly support the view thatregulation of the key enzyme of alcoholic fermentation, pyruvatedecarboxylase (PDC), is achieved in a number of ways, all associatedwith possible lowering of the cytoplasmic pH during anoxia.These mechanisms include not only the well-known acid pH optimumof PDC, but also long-term, reversible changes in characteristicsof the enzyme established both in vitro and in vivo. Following transfer of desalted extracts from pH 6.0 to 7.4,maximal activity of PDC was decreased, while there was a considerableincrease in the lag before maximal activity was reached. Similarchanges in enzyme characteristics were observed when wheat (Triticumaestivum L. cv. Gamenya) roots and rice (Oryza sativa L. cv.Calrose) coleoptiles were transferred from anoxic to aerobicsolutions, provided PDC was assayed within 10 min of the startof maceration. All of the above changes were usually readilyreversible when extracts were returned to pH 6.0, or when plantswere returned to anoxic solutions. Additional regulation of PDC would be achieved by the S0.5 forpyruvate which is 0.75 mol m–3 at pH 6.0, 1.0 mol m–3at pH 6.8, and 2.5 mol m–3 at pH 7.4; the latter is wellabove estimates for pyruvate concentrations in the cytoplasmof aerated tissues. We assess that the combined effects of the acid pH optimum,the high S0.5 at pH 7.4 and the long-term decreases in activityobserved during incubation at pH 7.4 would reduce PDC activityin aerobic cells to at most 7% of the activity in anoxic cells.Possible additional controls for the pathway of alcoholic fermentationare briefly considered. Key words: PDC, regulation, anoxia  相似文献   

12.
The monophenolase activity of Terfezia claveryi tyrosinase (EC 1.14.18.1) is described for the first time. This enzyme is fully latent and can only be detected if SDS is present in the reaction medium. Monophenolase activity was localized within the ascocarp using histochemical techniques. A detailed kinetic study of the parameters affecting this activity has been carried out. Both the characteristic lag period and the steady-state rate are affected by pH and the enzyme and substrate concentrations. The presence of catalytic concentrations of o -diphenols affected the lag period but not the steady-state rate. By increasing the concentration of o- diphenols, it was possible to evaluate the enzyme activation constant, Kact, which showed a value of 7.2 μ M . The experimental results are compatible with the mechanism previously described for tyrosinases from other sources.  相似文献   

13.
To obtain information on the biological significance of yeast fructose-2,6-bisphosphate 6-phosphatase, kinetic data of the purified enzyme [(1987) Eur. J. Biochem. 164, 27-30] have been measured. Maximal activity was found between pH 6 and 7, the apparent Michaelis constant with fructose 2,6-bisphosphate was 7.2 microM at pH 6.0 and 79 microM at pH 7.0. Concentrations required for 50% inhibition of the enzyme at pH 6.0 were 8 microM Fru2P, 45 microM G1c6P, 80 microM Fru6P and 200 microM inorganic phosphate. The known intracellular steady-state level of about 10 microM fructose 2,6-bisphosphate in the presence of glucose is likely to be the result of a balance between the rapid synthesis of fructose 2,6-bisphosphate catalyzed by 6-phosphofructose 2-kinase and a fructose 2,6-bisphosphate degrading activity. The biological function of fructose-2,6-bisphosphate 6-phosphatase with an apparent Michaelis constant between 7 and 79 microM fructose 2,6-bisphosphate at pH 6-7 is therefore suggested to participate in the maintenance of a steady-state level of fructose 2,6-bisphosphate in a concentration range that fits well with the Michaelis constant of the enzyme.  相似文献   

14.
At an alkaline pH and in an aqueous solution carbaryl hydrolyses to form 1-naphthol, methylamine and carbon dioxide, but it is much more stable at an acid pH. Soil perfusion column experiments indicated that the rate of carbaryl degradation at pH 6.0 to 7.0 was limited by the rate of chemical hydrolysis. Bacterial communities of at least 12 and 14 members were selected in continuous cultures using carbaryl as the sole carbon and nitrogen source at pH 6.0. These communities were supported by the slow formation of hydrolysis products and a carbaryl-degrading bacterium was not selected after greater than 2000 h. A bacterial community of at least eight members was selected using equimolar 1-naphthol and methylamine as its sole carbon and nitrogen source. In contrast, after a lag of between 10 and 50 days, soil perfusion column and continuous culture enrichments at pH 5.2 and 5.0, respectively, led to the selection of a Pseudomonas sp. which could utilize carbaryl as its sole carbon and nitrogen source.  相似文献   

15.
At an alkaline pH and in an aqueous solution carbaryl hydrolyses to form 1-naphthol, methylamine and carbon dioxide, but it is much more stable at an acid pH. Soil perfusion column experiments indicated that the rate of carbaryl degradation at pH 6.0 to 7.0 was limited by the rate of chemical hydrolysis. Bacterial communities of at least 12 and 14 members were selected in continuous cultures using carbaryl as the sole carbon and nitrogen source at pH 6.0. These communities were supported by the slow formation of hydrolysis products and a carbaryl-degrading bacterium was not selected after > 2000 h. A bacterial community of at least eight members was selected using equimolar 1-naphthol and methylamine as its sole carbon and nitrogen source. In contrast, after a lag of between 10 and 50 days, soil perfusion column and continuous culture enrichments at pH 5.2 and 5.0, respectively, led to the selection of a Pseudomonas sp. which could utilize carbaryl as its sole carbon and nitrogen source.  相似文献   

16.
A cytochrome oxidase was purified 52-fold from membranes of alkalophilic Bacillus firmus RAB by extraction with Triton X-100, ion-exchange and hydroxyapatite chromatography, and gel filtration. On denaturing gels, the purified enzyme dissociated into two subunits of 56,000 and 40,000 Mr as well as a cytochrome c with an Mr of approximately 14,000. Heme contents calculated for an enzyme with a molecular weight of 110,000 were found to be 2 mol of heme a and 1 mol of heme c per mol of cytochrome oxidase; approximately 2 mol of copper per mol of purified enzyme was also found. Enzyme activity was observed in assays using reduced yeast or horse heart cytochrome c. Activity of the purified enzyme was optimal at pH 6.0 and in the presence of added lipids. Impure, membrane-associated activity exhibited a broader pH range for optimal activity extending to alkaline values.  相似文献   

17.
The reaction between ligninase and hydrogen peroxide yielding Compound I has been investigated using a stopped-flow rapid-scan spectrophotometer. The optical absorption spectrum of Compound I appears different to that reported by Andrawis, A. et al. (1987) and Renganathan, V. and Gold, M.H. (1986), in that the Soret-maximum is at 401 nm rather than 408 nm. The second-order rate constant (4.2·105 M−1·s−1) for the formation of Compound I was independent of pH (pH 3.0–6.0). In the absence of external electron donors, Compound I decayed to Compound II with a half-life of 5–10 s at pH 3.1. The rate of this reaction was not affected by the H2O2 concentration used. In the presence of either veratryl alcohol or ferrocyanide, Compound II was rapidly generated. With ferrocyanide, the second-order rate constant increased from 1.9·104 M−1·s−1 to 6.8·106 M−1·s−1 when the pH was lowered from 6.0 to 3.1. With veratryl alcohol as an electron donor, the second-order rate constant for the formation of Compound II increased from 7.0·103 M−1·s−1 at pH 6.0 to 1.0·105 M−1·s−1 at pH 4.5. At lower pH values the rate of Compound II formation no longer followed an exponential relationship and the steady-state spectral properties differed to those recorded in the presence of ferrocyanide. Our data support a model of enzyme catalysis in which veratryl alcohol is oxidized in one-electron steps and strengthen the view that veratryl alcohol oxidation involves a substrate-modified Compound II intermediate which is rapidly reduced to the native enzyme.  相似文献   

18.
The binding stoichiometry of tRNATrp and tryptophanyl-tRNA synthetase (EC 6.1.1.2) from beef is examined by three approaches, under pH conditions of maximum activity (pH 8.0). (1) Analytical ultracentrifugation evidences the binding of a single mol of tRNATrp in a 2.5-10 microM concentration range. (2) tRNATrp quenches the fluorescence of the enzyme. The dependence of this fluorescence quenching on the tRNATrp concentration (0.1-4 microM) reflects also the binding of 1 mol of tRNA per mol of enzyme, with a Kd value of 0.19 +/- 0.02 microM. (3) tRNATrp protects the enzyme against derivatization by oxidized ATP. Out of the two fast-reacting lysine residues of the native enzyme, only one is prevented from reacting by tRNATrp in the 0.5-110 microM concentration range. This protection can be significantly analyzed only by assuming a one-to-one complex between the enzyme and tRNA. These results, obtained at pH 8.0 and 25 degrees C, are in contrast with the stoichiometry of 2 mol of tRNA to 1 mol of enzyme, previously observed at pH 6.0 and 4 degrees C.  相似文献   

19.
Modes of DNA cleavage by the EcoRV restriction endonuclease   总被引:6,自引:0,他引:6  
S E Halford  A J Goodall 《Biochemistry》1988,27(5):1771-1777
The mechanism of action of the EcoRV restriction endonuclease at its single recognition site on the plasmid pAT153 was analyzed by kinetic methods. In reactions at pH 7.5, close to the optimum for this enzyme, both strands of the DNA were cut in a single concerted reaction: DNA cut in only one strand of the duplex was neither liberated from the enzyme during the catalytic turnover nor accumulated as a steady-state intermediate. In contrast, reactions at pH 6.0 involved the sequential cutting of the two strands of the DNA. Under these conditions, DNA cut in a single strand was an obligatory intermediate in the reaction pathway and a fraction of the nicked DNA dissociated from the enzyme during the turnover. The different reaction profiles are shown to be consistent with a single mechanism in which the kinetic activity of each subunit of the dimeric protein is governed by its affinity for Mg2+ ions. At pH 7.5, Mg2+ is bound to both subunits of the dimer for virtually the complete period of the catalytic turnover, while at pH 6.0 Mg2+ is bound transiently to one subunit at a time. The kinetics of the EcoRV nuclease were unaffected by DNA supercoiling.  相似文献   

20.
M G Mulkerrin  R Wetzel 《Biochemistry》1989,28(16):6556-6561
Heated at pH 6.0 and at 50 degrees C, human interferon gamma (HuIFN-gamma) is inactivated via the formation of insoluble aggregates. At pH 6.0, the aggregation rate increases with temperature from 40 to 65 degrees C. There is a temperature-dependent time lag to aggregate formation observed in the generation of light-scattering particles at pH 6.0, and this correlates with the fast phase observed in the kinetics of reversible thermal unfolding. In addition, the dependence of aggregation kinetics on temperature closely follows the reversible melting curve. These observations suggest that at pH 6.0 irreversible thermal denaturation and aggregation depend on partial or complete unfolding of the molecule. At pH 5.0, also at 50 degrees C, the molecule is stable to irreversible aggregation. In reversible unfolding in 0.25 M guanidine hydrochloride, the Tm for HuIFN-gamma increases from 30.5 degrees C at pH 4.75 to 41.8 degrees C at pH 6.25, in analogy to the behavior of other globular proteins. These observations suggest that the relative instability of HuIFN-gamma to irreversible denaturation via aggregation at pH 6.0 compared to pH 5.0 is not due to an increased stability toward unfolding at the lower pH. Alternatively, stability at pH 5.0 must be due either to the improved solution properties of the unfolded state or to the improved solubility/decreased kinetic lifetime of an unfolding intermediate. Aggregation of HuIFN-gamma at 50 degrees C is half-maximal at pH 5.7, suggesting that protonation of one or both of the histidine residues may be involved in this stabilization.(ABSTRACT TRUNCATED AT 250 WORDS)  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