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.
J A Biosca  T E Barman  F Travers 《Biochemistry》1984,23(11):2428-2436
The initial steps by which ATP dissociates and binds to actomyosin subfragment 1 (acto-SF-1) were studied. Two techniques were used: stopped-flow (for acto-SF-1 dissociation kinetics) and rapid-flow quench with ATP chase quenching (for ATP binding kinetics). The experiments were carried out in 40% ethylene glycol-5 mM KCI, pH 8, at 15 degrees C. Under these conditions, the binding of SF-1 to actin remains very tight. As with SF-1, the ATP chase technique could be used, first, to titrate active sites and, second, to study the kinetics of ATP binding to acto-SF-1. The kinetic constants obtained were compared with those of SF-1 alone and with the acto-SF-1 dissociation kinetics under identical conditions. The kinetics of the acto-SF-1 dissociation did not vary with the actin to SF-1 ratio, but the ATP binding kinetics did, and a maximum value was reached at a mole ratio of 2.5. At high ATP (100 microM), kdiss = 300 s-1, which compares with 49 s-1 and 13 s-1 for the ATP binding kinetics for acto-SF-1 (actin to SF-1 = 1:1) and SF-1, respectively. As with SF-1, the ATP binding to acto-SF-1 follows a hyperbolic law with the ATP concentration. This suggests a rapid equilibrium (K) followed by an essentially irreversible step (k).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The initial steps of the ATPase of covalently cross-linked actomyosin subfragment 1 (acto-SF-1) were studied by the rapid flow quench method, and the results obtained were compared with those with reversible (i.e., non-cross-linked) acto-SF-1 and SF-1 under identical conditions. Cross-linked acto-SF-1 plus [gamma-32P]ATP reaction mixture milliseconds old was quenched either in a large excess of unlabeled ATP (ATP chase) or in acid (Pi burst). The conditions were pH 8 and 15 degrees C at 5 mM or 0.15 M KCl and with or without 40% ethylene glycol. In 40% ethylene glycol (5 mM KCl), as with SF-1 and reversible acto-SF-1, the ATP chase was used to titrate active sites and to study the kinetics of ATP binding. Unlike those with SF-1 or reversible acto-SF-1, saturation kinetics were not obtained. The second-order rate constant for ATP binding was 3.1 X 10(6) M-1 s-1 for cross-linked acto-SF-1, 1.8 X 10(6) M-1 s-1 for reversible acto-SF-1, and 2 X 10(6) M-1 s-1 for SF-1. In Pi burst experiments, a transient phase could not be discerned. Because of a high kcat, cross-linked acto-SF-1 was difficult to study in aqueous solution, but at 5 mM KCl, the ATP chase and Pi burst curves were similar to those obtained in 40% ethylene glycol. At 0.15 M KCl the ATP chase curve was difficult to interpret (small amplitude), and there was a small Pi burst.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
We have used a new fluorescent ATP analogue, 3'-(7-diethylaminocoumarin-3-carbonylamino)-3'-deoxyadenosine-5'-triphosphate (deac-aminoATP), to study the ATP hydrolysis mechanism of the single headed myosinV-S1. Our study demonstrates that deac-aminoATP is an excellent substrate for these studies. Although the deac-amino nucleotides have a low quantum yield in free solution, there is a very large increase in fluorescence emission ( approximately 20-fold) upon binding to the myosinV active site. The fluorescence emission intensity is independent of the hydrolysis state of the nucleotide bound to myosinV-S1. The very good signal-to-noise ratio that is obtained with deac-amino nucleotides makes them excellent substrates for studying expressed proteins that can only be isolated in small quantities. The combination of the fast rate of binding and the favorable signal-to-noise ratio also allows deac-nucleotides to be used in chase experiments to determine the kinetics of ADP and Pi dissociation from actomyosin-ADP-Pi. Although phosphate dissociation from actomyosinV-ADP-Pi does not itself produce a fluorescence signal, it produces a lag in the signal for deac-aminoADP dissociation. The lag provides direct evidence that the principal pathway of product dissociation from actomyosinV-ADP-Pi is an ordered mechanism in which phosphate precedes ADP. Although the mechanism of hydrolysis of deac-aminoATP by (acto)myosinV-S1 is qualitatively similar to the ATP hydrolysis mechanism, there are significant differences in some of the rate constants. Deac-aminoATP binds 3-fold faster to myosinV-S1, and the rate of deac-aminoADP dissociation from actomyosinV-S1 is 20-fold slower. Deac-aminoATP supports motility by myosinV-HMM on actin at a rate consistent with the slower rate of deac-aminoADP dissociation.  相似文献   

4.
Transient kinetic data of ATP binding and cleavage by cardiac myosin subfragment 1 (S1) were obtained by fluorescence stopped flow and analyzed by using computer modeling based on a consecutive, reversible two-step mechanism: (formula: see text) where M1 and M12 denote myosin species with enhanced fluorescence and K'O = K0/(K0[ATP] + 1). The kinetic constants K0, k12, k23, and k32 and the fractional contributions of M1 and M12 to the total fluorescence are analyzed over a range of systematically varied solution parameters. The initial ATP binding equilibrium (K0), which decreases with increasing pH, is facilitated by a positively charged protein residue with a pK of 7.1. An active-site charge of +1.5 is determined from the ionic strength dependence. The rate constants k12, k23, and k32 also exhibit pK's near neutrality but increase with increasing pH. The majority of the large (-54 kJ/mol) negative free energy of ATP binding occurs upon S1 isomerization, k12, and a large increase in entropy (183 J/kmol at 15 degrees C) is associated with the cleavage step. The equilibrium constant for the cleavage step, K2, is determined as 3.5 at pH 7.0, 15 degrees C, and 200 mM ionic strength. There are no significant changes in fractional contributions to total fluorescence enhancement due to solvent-dependent conformational changes of S1 in these data. When values for the combined rate constants are calculated and compared with those determined by graphical analysis, it is observed that graphical analysis overestimates the binding rate constant (K0k12) by 25% and the hydrolysis rate constant (k23 + k32) by as much as 30%.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Foster KA  Gilbert SP 《Biochemistry》2000,39(7):1784-1791
Ncd is a kinesin-related motor protein which drives movement to the minus-end of microtubules. The kinetics of Ncd were investigated using the dimeric construct MC1 (Leu(209)-Lys(700)) expressed in Escherichia coli strain BL21(DE) as a nonfusion protein [Chandra, R., Salmon, E. D., Erickson, H. P., Lockhart, A., and Endow, S. A. (1993) J. Biol. Chem. 268, 9005-9013]. Acid chemical quench flow methods were used to measure directly the rate of ATP hydrolysis, and stopped-flow kinetic methods were used to determine the kinetics of mantATP binding, mantADP release, dissociation of MC1 from the microtubule, and binding of MC1 to the microtubule. The results define a minimal kinetic mechanism, M.N + ATP M.N.ATP M.N.ADP.P N. ADP.P N.ADP + P M.N.ADP M.N + ADP, where N, M, and P represent Ncd, microtubules, and inorganic phosphate respectively, with k(+1) = 2.3 microM(-1) s(-1), k(+2) =23 s(-1), k(+3) =13 s(-1), k(+5)= 0.7 microM(-)(1) s(-)(1), and k(+6) = 3.7 s(-)(1). Phosphate release (k(+4)) was not measured directly although it is assumed to be fast relative to ADP release because Ncd is purified with ADP tightly bound at the active site. ATP hydrolysis occurs at 23 s(-)(1) prior to Ncd dissociation at 13 s(-)(1). The pathway for ATP-promoted detachment (steps 1-3) of Ncd from the microtubule is comparable to kinesin's. However, there are two major differences between the mechanisms of Ncd and kinesin. In contrast to kinesin, mantADP release for Ncd at 3.7 s(-)(1) is the slowest step in the pathway and is believed to limit steady-state turnover. Additionally, the burst amplitude observed in the pre-steady-state acid quench experiments is stoichiometric, indicating that Ncd, in contrast to kinesin, is not processive for ATP hydrolysis.  相似文献   

6.
Xie L  Li WX  Rhodes T  White H  Schoenberg M 《Biochemistry》1999,38(18):5925-5931
Alkylation of myosin's Cys-707 (SH1) and Cys-697 (SH2) has profound consequences for myosin's ability to interact with actin and hydrolyze MgATP. Pre-steady-state measurements of myosin-S1 alkylated at SH1 and SH2 by N-phenylmaleimide (NPM) in the presence of ATP were taken to identify the steps of the reaction that are altered. It was found that the rate constant most affected by this modification is the apparent rate of the ATP hydrolysis step. This rate constant is reduced 20000-fold, an effect comparable in magnitude to the effect of the same modification on the binding of MgATP to S1 or acto-S1 [Xie, L., and Schoenberg, M. (1998) Biochemistry 37, 8048]. In contrast, the rate constants of phosphate release and dissociation of acto-S1 by ATP were reduced <20-fold. For unmodified S1, the enhancement of fluorescence seen after addition of ATP had the same rate constant as the ATP hydrolysis step (S1.ATP if S1.ADP.Pi) measured by single-turnover experiments in a quench-flow experiment. This is consistent with results previously observed [Johnson, K. A., and Taylor, E. W. (1978) Biochemistry 17, 3432]. However, NPM-modified S1 exhibited virtually no fluorescence enhancement upon ATP binding. This provides further evidence that M.ATP is the predominant intermediate of NPM-S1-catalyzed ATP hydrolysis.  相似文献   

7.
We have recently proposed a "processive clamp" model for the ATP hydrolysis cycle of the nucleotide-binding domain (NBD) of the mitochondrial ABC transporter Mdl1 (Janas, E., Hofacker, M., Chen, M., Gompf, S., van der Does, C., and Tampé, R. (2003) J. Biol. Chem. 278, 26862-26869). In this model, ATP binding to two monomeric NBDs leads to formation of an NBD dimer that, after hydrolysis of both ATPs, dissociates and releases ADP. Here, we set out to follow the association and dissociation of NBDs using a novel minimally invasive site-specific labeling technique, which provides stable and stoichiometric attachment of fluorophores. The association and dissociation kinetics of the E599Q-NBD dimer upon addition and removal of ATP were determined by fluorescence self-quenching. Remarkably, the rate of ATP hydrolysis of the wild type NBD is determined by the rate of NBD dimerization. In the E599QNBD, however, in which the ATP hydrolysis is 250-fold reduced, the ATP hydrolysis reaction controls dimer dissociation and the overall ATPase cycle. These data explain contradicting observations on the rate-limiting step of various ABC proteins and further demonstrate that dimer formation is an important step in the ATP hydrolysis cycle.  相似文献   

8.
Phosphorylation of shark rectal Na,K-ATPase by ATP in the presence of Na(+) was characterized by chemical quench experiments and by stopped-flow RH421 fluorescence. The appearance of acid-stable phosphoenzyme was faster than the rate of fluorescence increase, suggesting that of the two acid-stable phosphoenzymes formed, RH421 exclusively detects formation of E(2)-P, which follows formation of E(1)-P. The stopped-flow RH421 fluorescence response to ATP phosphorylation was biphasic, with a major fast phase with k(obs) approximately 90 s(-1) and a minor slow phase with a k(obs) of approximately 9 s(-1) (20 degrees C, pH 7.4). The observed rate constants for both the slow and the fast phase could be fitted with identical second-degree functions of the ATP concentration with apparent binding constants of approximately 3.1 x 10(7) M(-1) and 1. 8 x 10(5) M(-1), respectively. Increasing [ADP] decreased k(obs) for the rate of the RH421 fluorescence response to ATP phosphorylation. This could be accounted for by the reaction of ADP with the initially formed E(1)-P followed by a conformational change to E(2)-P. The biphasic stopped-flow RH421 responses to ATP phosphorylation could be simulated, assuming that in the absence of K(+) the highly fluorescent E(2)-P is slowly transformed into the "K(+)-insensitive" E'(2)-P subconformation forming a side branch of the main cycle.  相似文献   

9.
The kinetics of ATP binding and hydrolysis (formation of acid-labile phosphate) by the Tetrahymena 30 S dynein ATPase has been measured by chemical quench flow methods. The amplitude of the ATP-binding transient gave a molecular weight per ATP-binding site of approximately 750,000, suggesting nearly 3 ATP binding sites/2 million Mr dynein molecule (Johnson, K. A., and Wall, J.S. (1983) J. Cell Biol. 96, 669-678). ATP binding occurred at the rate predicted from the apparent second order rate constant of 4.7 X 10(6) M-1 S-1 measured by analysis of the ATP-induced dissociation of the microtubule-dynein complex (Porter, M. E., and Johnson, K. A. (1983) J. Biol. Chem. 258, 6582-6587). Hydrolysis was slower than binding and occurred at a rate of 55 S-1, at 30 and 50 microM ATP. The rate limiting step for steady state turnover (product release) occurred with a rate constant of 8 S-1. These data show that the first two steps of the pathway of coupling ATP hydrolysis to the microtubule-dynein cross-bridge cycle are the same as those described by Lymn and Taylor for actomyosin (Lymn, R. W., and Taylor, E. W. (1971) Biochemistry 10, 4617-4624). Namely, ATP binding induces the very rapid dissociation of dynein from the microtubule and ATP hydrolysis occurs more slowly following dissociation. Moreover, in spite of rather gross structural differences, the kinetic constants for dynein and myosin are quite similar.  相似文献   

10.
Patterson-Ward J  Huang J  Lee I 《Biochemistry》2007,46(47):13593-13605
Lon is an ATP dependent serine protease responsible for degrading denatured, oxidatively damaged and certain regulatory proteins in the cell. In this study we exploited the fluorescence properties of a dansylated peptide substrate (S4) and the intrinsic Trp residues in Lon to monitor peptide interacting with the enzyme. We generated two proteolytically inactive Lon mutants, S679A and S679W, where the active site serine is mutated to an Ala and Trp residue, respectively. Stopped-flow fluorescence spectroscopy was used to identify key enzyme intermediates generated along the reaction pathway prior to peptide hydrolysis. A two-step peptide binding event is detected in both mutants, where a conformational change occurs after a rapid equilibrium peptide binding step. The Kd for the initial peptide binding step determined by kinetic and equilibrium binding techniques is approximately 164 micromolar and 38 micromolar, respectively. The rate constants for the conformational change detected in the S679A and S679W Lon mutants are 0.74 +/- 0.10 s(-1) and 0.57 +/- 0.10 s(-1), respectively. These values are comparable to the lag rate constant determined for peptide hydrolysis (klag approximately 1 s(-1)) [Vineyard, D., et al. (2005) Biochemistry 45, 4602-4610]. Replacement of the active site Ser with Trp (S679W) allows for the detection of an ATP-dependent conformational change within the proteolytic site. The rate constant for this conformational change is 7.6 +/- 1.0 s(-1), and is essentially identical to the burst rate constant determined for ATP hydrolysis under comparable reaction conditions. Collectively, these kinetic data support a mechanism by which the binding of ATP to an allosteric site on Lon activates the proteolytic site. In this model, the energy derived from the binding of ATP minimally supports peptide cleavage by allowing peptide substrate access to the proteolytic site. However, the kinetics of peptide cleavage are enhanced by the hydrolysis of ATP.  相似文献   

11.
The fluorescent nucleotides epsilon ADP and epsilon ATP were used to study the binding and hydrolysis mechanisms of subfragment 1 (S-1) and acto-subfragment 1 from striated and smooth muscle. The quenching of the enhanced fluorescence emission of bound nucleotide by acrylamide analyzed either by the Stern-Volmer method or by fluorescence lifetime measurements showed the presence of two bound nucleotide states for 1-N6-ethenoadenosine triphosphate (epsilon ATP), 1-N6-ethenoadenosine diphosphate (epsilon ADP), and epsilon ADP-vanadate complexes with S-1. The equilibrium constant relating the two bound nucleotide states was close to unity. Transient kinetic studies showed two first-order transitions with rate constants of approximately 500 and 100 s-1 for both epsilon ATP and epsilon ADP and striated muscle S-1 and 300 and 30 s-1, respectively, for smooth muscle S-1. The hydrolysis of [gamma-32P] epsilon ATP yielded a transient phase of small amplitude (less than 0.2 mol/site) with a rate constant of 5-10 s-1. Consequently, the hydrolysis of the substrate is a step in the mechanism which is distinct from the two conformational changes induced by the binding of epsilon ATP. An essentially symmetric reaction mechanism is proposed in which two structural changes accompany substrate binding and the reversal of these steps occurs in product release. epsilon ATP dissociates acto-S-1 as effectively as ATP. For smooth muscle acto-S-1, dissociation proceeds in two steps, each accompanied by enhancement of fluorescence emission. A symmetric reaction scheme is proposed for the acto-S-1 epsilon ATPase cycle. The very similar kinetic properties of the reactions of epsilon ATP and ATP with S-1 and acto-S-1 suggest that two ATP intermediate states also occur in the ATPase reaction mechanism.  相似文献   

12.
The kinetics of ATP-induced dissociation of dynein from the dynein-microtubule complex has been investigated by stopped flow light scattering methods. The addition of ATP to the dynein-microtubule complex induced a large, rapid decrease in light scattering followed by a smaller and much slower decrease. The fast light scattering change was shown to be a measure of the ATP-induced dissociation of dynein from the dynein-microtubule complex and was distinguished from microtubule disassembly by several criteria. (i) The fast reaction occurred over a period of milliseconds and the rate was a function of the ATP concentration, whereas, the slow reaction occurred over a period of several seconds and was independent of ATP concentration; (ii) the amplitude of the fast reaction was directly proportional to the amount of dynein bound to the microtubule lattice; and (iii) only the slow phase was inhibited by the addition of the microtubule-stabilizing drug, taxol. The rate of ATP-induced dissociation of dynein from the microtubule increased linearly with increasing ATP concentration to give an apparent second order rate constant for ATP binding equal to k1 = 4.7 X 10(6) M-1 s-1 according to the following pathway: (formula; see text) where M X D represents the dynein-microtubule complex and D represents dynein. The loss of signal amplitude at high ATP concentration provided a minimum estimate for the rate of dissociation of the ternary complex (M X D X ATP) equal to kd greater than 1000 s-1. Thus, the dynein-microtubule system is similar to actomyosin in that ATP induces an extremely rapid dissociation of dynein from the microtubule.  相似文献   

13.
Xia Z  Azurmendi HF  Mildvan AS 《Biochemistry》2005,44(46):15334-15344
The MutT pyrophosphohydrolase, in the presence of Mg2+, catalyzes the hydrolysis of nucleoside triphosphates by nucleophilic substitution at Pbeta, to yield the nucleotide and PP(i). The best substrate for MutT is the mutagenic 8-oxo-dGTP, on the basis of its Km being 540-fold lower than that of dGTP. Product inhibition studies have led to a proposed uni-bi-iso kinetic mechanism, in which PP(i) dissociates first from the enzyme-product complex (k3), followed by NMP (k4), leaving a product-binding form of the enzyme (F) which converts to the substrate-binding form (E) in a partially rate-limiting step (k5) [Saraswat, V., et al. (2002) Biochemistry 41, 15566-15577]. Single- and multiple-turnover kinetic studies of the hydrolysis of dGTP and 8-oxo-dGTP and global fitting of the data to this mechanism have yielded all of the nine rate constants. Consistent with an "iso" mechanism, single-turnover studies with dGTP and 8-oxo-dGTP hydrolysis showed slow apparent second-order rate constants for substrate binding similar to their kcat/Km values, but well below the diffusion limit (approximately 10(9) M(-1) s(-1)): k(on)app = 7.2 x 10(4) M(-1) s(-1) for dGTP and k(on)app = 2.8 x 10(7) M(-1) s(-1) for 8-oxo-dGTP. These low k(on)app values are fitted by assuming a slow iso step (k5 = 12.1 s(-1)) followed by fast rate constants for substrate binding: k1 = 1.9 x 10(6) M(-1) s(-1) for dGTP and k1 = 0.75 x 10(9) M(-1) s(-1) for 8-oxo-dGTP (the latter near the diffusion limit). With dGTP as the substrate, replacing Mg2+ with Mn2+ does not change k1, consistent with the formation of a second-sphere MutT-M2+-(H2O)-dGTP complex, but slows the iso step (k5) 5.8-fold, and its reverse (k(-5)) 25-fold, suggesting that the iso step involves a change in metal coordination, likely the dissociation of Glu-53 from the enzyme-bound metal so that it can function as the general base. Multiple-turnover studies with dGTP and 8-oxo-dGTP show bursts of product formation, indicating partially rate-limiting steps following the chemical step (k2). With dGTP, the slow steps are the chemical step (k2 = 10.7 s(-1)) and the iso step (k5 = 12.1 s(-1)). With 8-oxo-dGTP, the slow steps are the release of the 8-oxo-dGMP product (k4 = 3.9 s(-1)) and the iso step (k5 = 12.1 s(-1)), while the chemical step is fast (k2 = 32.3 s(-1)). The transient kinetic studies are generally consistent with the steady state kcat and Km values. Comparison of rate constants and free energy diagrams indicate that 8-oxo-dGTP, at low concentrations, is a better substrate than dGTP because it binds to MutT 395-fold faster, dissociates 46-fold slower, and has a 3.0-fold faster chemical step. The true dissociation constants (KD) of the substrates from the E-form of MutT, which can now be obtained from k(-1)/k1, are 3.5 nM for 8-oxo-dGTP and 62 microM for dGTP, indicating that 8-oxo-dGTP binds 1.8 x 10(4)-fold tighter than dGTP, corresponding to a 5.8 kcal/mol lower free energy of binding.  相似文献   

14.
Hsc66 from Escherichia coli is a constitutively expressed hsp70 class molecular chaperone whose activity is coupled to ATP binding and hydrolysis. To better understand the mechanism and regulation of Hsc66, we investigated the kinetics of ATP hydrolysis and the interactions of Hsc66 with nucleotides. Steady-state experiments revealed that Hsc66 has a low affinity for ATP (K(m)(ATP) = 12.7 microM) compared with other hsp70 chaperones. The kinetics of nucleotide binding were determined by analyzing changes in the Hsc66 absorbance spectrum using stopped-flow methods at 23 degrees C. ATP binding results in a rapid, biphasic increase of Hsc66 absorbance at 280 nm; this is interpreted as arising from a two-step process in which ATP binding (k(a)(ATP) = 4.2 x 10(4) M(-1) s(-1), k(d)(ATP) = 1.1 s(-1)) is followed by a slow conformational change (k(conf) = 0. 1 s(-1)). Under single turnover conditions, the ATP-induced transition decays exponentially with a rate (k(decay) = 0.0013 s(-1)) similar to that observed in both steady-state and single turnover ATP hydrolysis experiments (k(hyd) = 0.0014 s(-1)). ADP binding to Hsc66 results in a monophasic transition in the absence (k(a)(ADP) = 7 x 10(5) M(-1) s(-1), k(d)(ADP) = 60 s(-1)) and presence of physiological levels of inorganic phosphate (k(a)(ADP(P(i)) = 0.28 x 10(5) M(-1) s(-1), k(d)(ADP(P(i)) = 9.1 s(-1)). These results indicate that ATP hydrolysis is the rate-limiting step under steady-state conditions and is >10(3)-fold slower than the rate of ADP/ATP exchange. Thus, in contrast to DnaK and eukaryotic forms of hsp70 that have been characterized to date, the R if T equilibrium balance for Hsc66 is shifted in favor of the low peptide affinity T state, and regulation of the reaction cycle is expected to occur at the ATP hydrolysis step rather than at nucleotide exchange.  相似文献   

15.
ClpX requires ATP to unfold protein substrates and translocate them into the proteolytic chamber of ClpP for degradation. The steady-state parameters for hydrolysis of ATP and ATPgammaS by ClpX were measured with different protein partners and the kinetics of degradation of ssrA-tagged substrates were determined with both nucleotides. ClpX hydrolyzed ATPgammaS to ADP and thiophosphate at a rate (6/min) significantly slower than ATP hydrolysis (140/min), but the hydrolysis of both nucleotides was increased by ssrA-tagged substrates and decreased by ClpP. K(M) and k(cat) for hydrolysis of ATP and ATPgammaS were linearly correlated over a 200-fold range, suggesting that protein partners largely affect k(cat) rather than nucleotide binding, indicating that most bound ATP leaves the enzyme by hydrolysis rather than dissociation, and placing an upper limit of approximately 15 micro M on K(D) for both nucleotides. Competition studies with ClpX and fluorescently labeled ADP gave inhibition constants for ATPgammaS ( approximately 2 micro M) and ADP ( approximately 3 micro M) under the reaction conditions used for steady-state kinetics. In the absence of Mg(2+), where hydrolysis does not occur, the inhibition constant for ATP ( approximately 55 micro M) was weaker but very similar to the value for ATPgammaS ( approximately 45 micro M). Compared with ATP, ATPgammaS supported slow but roughly comparable rates of ClpXP degradation for two Arc-ssrA substrates and denatured GFP-ssrA, but not of native GFP-ssrA. These results show that the processing of protein substrates by ClpX is closely coupled to the maximum rate of nucleotide hydrolysis.  相似文献   

16.
The large change in fluorescence emission of 1-N6-etheno-2-aza-ATP (epsilon-aza-ATP) has been used to investigate the kinetic mechanism of etheno-aza nucleotide binding to bovine cardiac myosin subfragment 1 (myosin-S1) and actomyosin subfragment 1 (actomyosin-S1). The time course of nucleotide fluorescence enhancement observed during epsilon-aza-ATP hydrolysis is qualitatively similar to the time course of tryptophan fluorescence enhancement observed during ATP hydrolysis. In single turnover experiments, the nucleotide fluorescence rapidly increases to a maximum level, then decreases with a rate constant of 0.045 s-1 to a final level, which is about 30% of the maximal enhancement; a similar fluorescence enhancement is obtained by adding epsilon-aza-ADP to cardiac myosin-S1 or actomyosin-S1 under the same conditions (100 mM KCl, 10 mM 4-morpholinepropanesulfonic acid, 5 mM MgCl2, 0.1 mM dithiothreitol, pH 7.0, 15 degrees C). The kinetic data are consistent with a mechanism in which there are two sequential (acto)myosin-S1 nucleotide complexes with enhanced nucleotide fluorescence following epsilon-aza-ATP binding. The apparent second order rate constants of epsilon-aza-ATP binding to cardiac myosin subfragment 1 and actomyosin subfragment 1 are 2-12 times slower than those for ATP. Actin increases the rate of epsilon-aza-ADP dissociation from bovine cardiac myosin-S1 from 1.9 to 110 s-1 at 15 degrees C which can be compared to 0.3 and 65 s-1 for ADP dissociation under similar conditions. Although there are quantitative differences between the rate and equilibrium constants of epsilon-aza- and adenosine nucleotides to cardiac actomyosin-S1 and myosin-S1, the basic features of the nucleotide binding steps of the mechanism are unchanged.  相似文献   

17.
Although there is agreement that actomyosin can hydrolyze ATP without dissociation of the actin from myosin, there is still controversy about the nature of the rate-limiting step in the ATPase cycle. Two models, which differ in their rate-limiting step, can account for the kinetic data. In the four-state model, which has four states containing bound ATP or ADP . Pi, the rate-limiting step is ATP hydrolysis (A . M . ATP in equilibrium A . M . ADP . Pi). In the six-state model, which we previously proposed, the rate-limiting step is a conformational change which occurs before Pi release but after ATP hydrolysis. A difference between these models is that only the four-state model predicts that almost no acto-subfragment 1 (S-1) . ADP . Pi complex will be formed when ATP is mixed with acto . S-1. In the present study, we determined the amount of acto . S-1 . ADP . Pi formed when ATP is mixed with S-1 cross-linked to actin [Mornet, D., Bertrand, R., Pantel, P., Audemard, E., & Kassab, R. (1981) Nature (London) 292, 301-306]. The amount of acto . S-1 . ADP . Pi was determined both from intrinsic fluorescence enhancement and from direct measurement of Pi. We found that at mu = 0.013 M, the fluorescence magnitude in the presence of ATP of the cross-linked actin . S-1 preparation was about 50% of the value obtained with S-1, while at mu = 0.053 M the fluorescence magnitude was about 70% of that obtained with S-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
The sarcoplasmic calcium-binding protein (SCP) of the sandworm Nereis possesses three Ca2(+)-Mg2+ sites but no Ca2(+)-specific site. Binding of Mg2+, but not of Ca2+, displays a marked positive cooperativity. The apparent cooperativity of Ca2+ binding in the presence of Mg2+ results from the allostery in Mg2+ dissociation. Binding of the first Ca2+ or Mg2+ induces all the conformational change, monitored by Trp fluorescence. In displacement reactions the conformational changes occur in the step SCP.Mg3----SCP.Ca1Mg2. Stopped-flow experiments indicate that Trp fluorescence changes upon Ca2(+)-binding are instantaneous whereas Mg2(+)-binding involves a fast pre-equilibrium (Keq = 28 M-1), followed by two slow consecutive conformational changes with k1 = 13.5 s-1 and k2 = 0.21 s-1. The fluorescence change after dissociation of Ca2+ from SCP is monophasic with k = 0.02 s-1; that after Mg2+ dissociation is biphasic with k1 = 0.8 s-1 and k2 = 0.1 s-1. Trp life time measurements also indicate that Ca2(+)- and Mg2(+)-induced conformational changes are completely different. Displacement of bound Ca2+ by Mg2+ can be described by two consecutive reactions in which the first (without fluorescence change) corresponds to the dissociation of the last Ca2+ (k1 = 2.4 s-1) and the second (k2 = 0.45 s-1) to the final conformational change observed upon direct Mg2+ binding. Displacement of bound Mg2+ by Ca2+ follows the kinetic scheme of simple competition; the conformational rate constant approaches asymptotically (up to the limit of 129 s-1) the dissociation rate of Mg2+ as the concentration of Ca2+ increases. In summary, after fast dissociation of Ca2+ or Mg2+, Nereis SCP slowly converts to the metal-free configuration, but in Ca2(+)-Mg2+ exchange reactions, the conformational changes are nearly as fast as the cation dissociation reactions.  相似文献   

19.
S K Singh  F Guo  M R Maurizi 《Biochemistry》1999,38(45):14906-14915
The Escherichia coli ClpA and ClpP proteins form a complex, ClpAP, that catalyzes ATP-dependent degradation of proteins. Formation of stable ClpA hexamers and stable ClpAP complexes requires binding of ATP or nonhydrolyzable ATP analogues to ClpA. To understand the order of events during substrate binding, unfolding, and degradation by ClpAP, it is essential to know the oligomeric state of the enzyme during multiple catalytic cycles. Using inactive forms of ClpA or ClpP as traps for dissociated species, we measured the rates of dissociation of ClpA hexamers or ClpAP complexes. When ATP was saturating, the rate constant for dissociation of ClpA hexamers was 0.032 min(-1) (t(1/2) of 22 min) at 37 degrees C, and dissociation of ClpP from the ClpAP complexes occurred with a rate constant of 0. 092 min(-1) (t(1/2) of 7.5 min). Because the k(cat) for casein degradation is approximately 10 min(-1), these results indicate that tens of molecules of casein can be turned over by the ClpAP complex before significant dissociation occurs. Mutations in the N-terminal ATP binding site led to faster rates of ClpA and ClpAP dissociation, whereas mutations in the C-terminal ATP binding site, which cause significant decreases in ATPase activity, led to lower rates of dissociation of ClpA and ClpAP complexes. Dissociation rates for wild-type and first domain mutants of ClpA were faster at low nucleotide concentrations. The t(1/2) for dissociation of ClpAP complexes in the presence of nonhydrolyzable analogues was >/=30 min. Thus, ATP binding stabilizes the oligomeric state of ClpA, and cycles of ATP hydrolysis affect the dynamics of oligomer interaction. However, since the k(cat) for ATP hydrolysis is approximately 140 min(-1), ClpA and the ClpAP complex remain associated during hundreds of rounds of ATP hydrolysis. Our results indicate that the ClpAP complex is the functional form of the protease and as such engages in multiple rounds of interaction with substrate proteins, degradation, and release of peptide products without dissociation.  相似文献   

20.
The mechanism of kinesin ATPase has been investigated by transient state kinetic analysis. The results satisfy the scheme [formula: see text] where T, D, and P(i) refer to nucleotide tri- and diphosphate and inorganic phosphate, respectively. The nucleotide-binding steps were measured by the fluorescence enhancement of mant (2'-(3')-O-(N-methylanthraniloyl)-ATP and mant-ADP. The initial rapid equilibrium binding steps (1) and (6) are followed by isomerizations (k2 = 170 +/- 30 s-1 at 20 degrees C, k-5 greater than 100 s-1). The increase in fluorescence is 20-25% larger for K.T** than K.D*. The rate constant of the hydrolysis step k3 is 6-7 s-1. The fluorescence decreases after formation of K.T** at a rate of 7-10 s-1. This change could occur in step 3 or in step 4 if k4 much greater than k3. The value of k4 is larger than 0.1 s-1. The steady state rate is 0.003 s-1 which agrees with the rate of ADP dissociation (k5). Step 5 is rate limiting in the scheme in agreement with the conclusion of Hackney (Hackney, D. D. (1988) Proc. Natl. Acad. Sci. U.S.A. 85, 6314-6318) that ADP dissociation is the rate-limiting step.  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