首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 78 毫秒
1.
We have studied the kinetics of "unisite" ATP hydrolysis and synthesis in seven mutant Escherichia coli F1-ATPase enzymes. The seven mutations are distributed over a 105-residue segment of the catalytic nucleotide-binding domain in beta-subunit and are: G142S, K155Q, K155E, E181Q, E192Q, M209I, and R246C. We report forward and reverse rate constants and equilibrium constants in all seven mutant enzymes for the four steps of unisite kinetics, namely (i) ATP binding/release, (ii) ATP hydrolysis/synthesis, (iii) Pi release/binding, and (iv) ADP release/binding. The seven mutant enzymes displayed a wide range of deviations from normal in both rate and equilibrium constants, with no discernible common pattern. Notably, steep reductions in Kd ATP were seen in some cases, the value of Kd Pi was high, and K2 (ATP hydrolysis/synthesis) was relatively unaffected. Significantly, when the data from the seven mutations were combined with previous data from two other E. coli F1-beta-subunit mutations (D242N, D242V), normal E. coli F1, soluble and membranous mitochondrial F1, it was found that linear free energy relationships obtained for both ATP binding/release (log k+1 versus log K1) and ADP binding/release (log k-4 versus log K-4). Two conclusions follow. 1) The seven mutations studied here cause subtle changes in interactions between the catalytic nucleotide-binding domain and substrate ATP or product ADP. 2) The mitochondrial, normal E. coli, and nine total beta-subunit mutant enzymes represent a continuum in which subtle structural differences in the catalytic site resulted in changes in binding energy; therefore insights into the nature of energy coupling during ATP hydrolysis and synthesis by F1-ATPase may be ascertained by detailed studies of this group of enzymes.  相似文献   

2.
Microtubules accelerate ADP release by dynein   总被引:4,自引:0,他引:4  
E L Holzbaur  K A Johnson 《Biochemistry》1989,28(17):7010-7016
The effects of microtubules on the phosphate-water oxygen exchange reactions catalyzed by dynein were examined in order to determine the mechanism by which microtubules activate the ATPase. Microtubules inhibited the rate of medium exchange observed during net ATP hydrolysis. Inhibition of the exchange reaction was proportional to the extent of microtubule activation of ATP turnover with no effect on the partition coefficient. These data argue that microtubules do not increase the rate of release of phosphate from dynein; rather, they increase the rate of ADP release. Microtubules markedly inhibited medium phosphate-water exchange reactions observed in the presence of ADP and Pi. With increasing concentrations of ADP, the rate of exchange increased in parallel to the dissociation of dynein from the microtubules, suggesting that only free dynein and not the microtubule-dynein complex catalyzes the exchange reaction. The rates of dynein binding to microtubules in the absence and presence of saturating ADP were 1.6 X 10(6) and 9.8 X 10(5) M-1 s-1, respectively. ADP inhibited the rate of the ATP-induced dissociation of the microtubule-dynein complex with an apparent Kd = 0.37 mM for the binding of ADP to the microtubule-dynein complex. However, the rate of dissociation of ADP from the M.D.ADP complex was quite fast (approximately 1000 s-1). These data support the postulate of a high-energy dynein-ADP intermediate and indicate that microtubules activate the dynein ATPase by enhancing the rate of ADP release.  相似文献   

3.
E L Holzbaur  K A Johnson 《Biochemistry》1989,28(13):5577-5585
The kinetics of the product release steps in the pathway of ATP hydrolysis by dynein were investigated by examining the rate and partition coefficient of phosphate-water 18O exchange under equilibrium and steady-state conditions. Dynein catalyzed both medium and intermediate phosphate-water oxygen exchange with a partition coefficient of 0.30. The dependence of the rate of loss of the fully labeled phosphate species on the concentration of ADP was hyperbolic, with an apparent Kd for the binding of ADP to dynein of 0.085 mM. The apparent second-order rate constant for phosphate binding to the dynein-ADP complex was 8000 M-1 s-1. The time course of medium phosphate-water oxygen exchange during net ATP hydrolysis was examined in the presence of an ATP regeneration system. The observed rate of loss of P18O4 was comparable to the rate observed at saturating ADP which implies that ADP release is rate limiting for dynein in the steady state. Product inhibition of the dynein ATPase was also examined. ADP inhibited the enzyme competitively with a Ki of 0.4 mM. Phosphate was a linear noncompetitive mixed-type inhibitor with a Ki of 11 mM. These data were fit to a model in which phosphate release is fast and is followed by rate-limiting release of ADP, allowing us to define each rate constant in the pathway. A discrepancy between the total free energy calculated compared to the known free energy of ATP hydrolysis suggests that there is an additional step in the pathway, perhaps involving a change in conformation of the enzyme-ADP state preceding ADP release.  相似文献   

4.
The oxygen isotope exchange reactions catalyzed by sea urchin Strongylocentrotus intermedius spermatozoa dynein I were studied with a view of comparing molecular mechanisms of ATP hydrolysis by dynein and myosin ATPases. It was demonstrated that the isotope exchange takes place during ATP hydrolysis and during enzyme incubation with ADP and Pi and is absent when the enzyme is incubated with Pi. It was assumed that the molecular mechanisms of ATP hydrolysis by dynein I and myosin are identical.  相似文献   

5.
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   

6.
The Mg2+-dependent ATPase (adenosine 5'-triphosphatase) mechanism of myosin and subfragment 1 prepared from frog leg muscle was investigated by transient kinetic technique. The results show that in general terms the mechanism is similar to that of the rabbit skeletal-muscle myosin ATPase. During subfragment-1 ATPase activity at 0-5 degrees C pH 7.0 and I0.15, the predominant component of the steady-state intermediate is a subfragment-1-products complex (E.ADP.Pi). Binary subfragment-1-ATP (E.ATP) and subfragment-1-ADP (E.ADP) complexes are the other main components of the steady-state intermediate, the relative concentrations of the three components E.ATP, E.ADP.Pi and E.ADP being 5.5:92.5:2.0 respectively. The frog myosin ATPase mechanism is distinguished from that of the rabbit at 0-5 degrees C by the low steady-state concentrations of E.ATP and E.ADP relative to that of E.ADP.Pi and can be described by: E + ATP k' + 1 in equilibrium k' - 1 E.ATP k' + 2 in equilibrium k' - 2 E.ADP.Pi k' + 3 in equilibrium k' - 3 E.ADP + Pi k' + 4 in equilibrium k' - 4 E + ADP. In the above conditions successive forward rate constants have values: k' + 1, 1.1 X 10(5)M-1.S-1; k' + 2 greater than 5s-1; k' + 3, 0.011 s-1; k' + 4, 0.5 s-1; k'-1 is probably less than 0.006s-1. The observed second-order rate constants of the association of actin to subfragment 1 and of ATP-induced dissociation of the actin-subfragment-1 complex are 5.5 X 10(4) M-1.S-1 and 7.4 X 10(5) M-1.S-1 respectively at 2-5 degrees C and pH 7.0. The physiological implications of these results are discussed.  相似文献   

7.
The actin-myosin lattice spacing of rabbit psoas fibers was osmotically compressed with a dextran T-500, and its effect on the elementary steps of the cross-bridge cycle was investigated. Experiments were performed at the saturating Ca (pCa 4.5-4.9), 200 mM ionic strength, pH 7.0, and at 20 degrees C, and the results were analyzed by the following cross-bridge scheme: [formula: see text] where A = actin, M = myosin head, S = MgATP, D = MgADP, and P = Pi = phosphate. From MgATP and MgADP studies on exponential process (C) and (D), the association constants of cross-bridges to MgADP (K0), MgATP (K1a), the rate constants of the isomerization of the AM S state (k1b and k-1b), and the rate constants of the cross-bridge detachment step (k2 and k-2) were deduced. From Pi study on process (B), the rate constants of the cross-bridge attachment (power stroke) step (k4- and k-4) and the association constant of Pi ions to cross-bridges (K5) were deduced. From ATP hydrolysis measurement, the rate constant of ADP-isomerization (rate-limiting) step (k6) was deduced. These kinetic constants were studied as functions of dextran concentrations. Our results show that nucleotide binding, the ATP-isomerization, and the cross-bridge detachment steps are minimally affected by the compression. The rate constant of the reverse power stroke step (k-4) decreases with mild compression (0-6.3% dextran), presumably because of the stabilization of the attached cross-bridges in the AM*DP state. The rate constant of the power stroke step (k4) does not change with mild compression, but it decreases with higher compression (> 6.3% dextran), presumably because of an increased difficulty in performing the power stroke. These results are consistent with the observation that isometric tension increases with a low level of compression and decreases with a high level of compression. Our results also show that the association constant K5 of Pi with cross-bridge state AM*D is not changed with compression. Our result further show that the ATP hydrolysis rate decreased with compression, and that the rate constants of the ADP-isomerization step (k6) becomes progressively less with compression. The effect of compression on the power stroke step and rate-limiting step implies that a large-scale molecular rearrangement in the myosin head takes place in these two slow reaction steps.  相似文献   

8.
Bicarbonate, an activating anion of ATP hydrolysis, inhibited ATP synthesis coupled to succinate oxidation in beef heart submitochondrial particles but diminished the lag time and increased the steady-state velocity of the (32)Pi-ATP exchange reaction. The latter effects exclude the possibility that bicarbonate is inducing an intrinsic uncoupling between ATP hydrolysis and proton translocation at the level of F(1)F(o) ATPase. The inhibition of ATP synthesis was competitive with respect to ADP at low fixed [Pi], mixed at high [Pi] and non-competitive towards Pi at any fixed [ADP]. From these results we can conclude that (i) bicarbonate does not bind to a Pi site in the mitochondrial F(1); (ii) it competes with the binding of ADP to a low-affinity site, likely the low-affinity non-catalytic nucleotide binding site. It is postulated that bicarbonate stimulates ATP hydrolysis and inhibits ATP synthesis by modulating the relative affinities of the catalytic site for ATP and ADP.  相似文献   

9.
Escherichia coli Rep helicase catalyzes the unwinding of duplex DNA in reactions that are coupled to ATP binding and hydrolysis. We have investigated the kinetic mechanism of ATP binding and hydrolysis by a proposed intermediate in Rep-catalyzed DNA unwinding, the Rep "P2S" dimer (formed with the single-stranded (ss) oligodeoxynucleotide, (dT)16), in which only one subunit of a Rep homo-dimer is bound to ssDNA. Pre-steady-state quenched-flow studies under both single turnover and multiple turnover conditions as well as fluorescence stopped-flow studies were used (4 degrees C, pH 7.5, 6 mM NaCl, 5 mM MgCl2, 10 % (v/v) glycerol). Although steady-state studies indicate that a single ATPase site dominates the kinetics (kcat=17(+/-2) s-1; KM=3 microM), pre-steady-state studies provide evidence for a two-ATP site mechanism in which both sites of the dimer are catalytically active and communicate allosterically. Single turnover ATPase studies indicate that ATP hydrolysis does not require the simultaneous binding of two ATP molecules, and under these conditions release of product (ADP-Pi) is preceded by a slow rate-limiting isomerization ( approximately 0.2 s-1). However, product (ADP or Pi) release is not rate-limiting under multiple turnover conditions, indicating the involvement of a second ATP site under conditions of excess ATP. Stopped-flow fluorescence studies monitoring ATP-induced changes in Rep's tryptophan fluorescence displayed biphasic time courses. The binding of the first ATP occurs by a two-step mechanism in which binding (k+1=1.5(+/-0.2)x10(7) M-1 s-1, k-1=29(+/-2) s-1) is followed by a protein conformational change (k+2=23(+/-3) s-1), monitored by an enhancement of Trp fluorescence. The second Trp fluorescence quenching phase is associated with binding of a second ATP. The first ATP appears to bind to the DNA-free subunit and hydrolysis induces a global conformational change to form a high energy intermediate state with tightly bound (ADP-Pi). Binding of the second ATP then leads to the steady-state ATP cycle. As proposed previously, the role of steady-state ATP hydrolysis by the DNA-bound Rep subunit may be to maintain the DNA-free subunit in an activated state in preparation for binding a second fragment of DNA as needed for translocation and/or DNA unwinding. We propose that the roles of the two ATP sites may alternate upon binding DNA to the second subunit of the Rep dimer during unwinding and translocation using a subunit switching mechanism.  相似文献   

10.
K A Johnson  E W Taylor 《Biochemistry》1978,17(17):3432-3442
The kinetics of the increase in protein fluorescence following the addition of ATP to subfragment-1 (SF-1) and acto-SF-1 have been reinvestigated. The concentration dependence of the rate obtained with SF-1 did not fit a hyperbola and at high ATP concentration, approximately 40% of the signal amplitude was lost due to a fast phase at the beginning of the transient (20 degrees C). At lower temperature (less than or equal to 10 degrees C) the fluorescence transient was biphasic, with a fast phase observed at high ATP concentration. These results indicate that there are two steps in the SF-1 pathway in which there is a change in protein fluorescence. Measurements of ATP binding and hydrolysis by chemical quench-flow methods indicate that the rate of ATP binding is correlated with the fast fluorescence step and hydrolysis is correlated with the slow fluorescence change. The SF-1 mechanism can thus be described as: (formula: see text) where M represents SF-1 and states of enhanced fluorescence are given by M (16%) and M (36% enhancement, relative to SF-1). Step 1 is a rapid equilibrium with K1 approximately 10(3) M-1. Tight binding of ATP occurs in step 2 and the loss of signal amplitude requires k2 greater than or approximately 1500--2000 s-1. The maximum observed fluorescence rate defines the rate of hydrolysis, k3 + k-3 = 125 s-1 (20 degrees C, 0.1 M KCl, pH 7.0). The steps in the mechanism correspond to the Bagshaw--Trentham scheme, with the important difference that the assignment of rate constant is altered. Formation of the acto-SF-1 complex gave a fluorescence enhancement of approximately 14% relative to SF-1. Dissociation of acto-SF-1 by ATP produced a 20--22% enhancement in fluorescence. There was no detectable fluorescence change during dissociation as evidenced by a lag in the fluorescence transient which corresponded to the kinetics of dissociation. The fluorescence change occurred at the same maximum rate as for SF-1 but there was no loss in signal amplitude at high ATP concentration. The kinetics of the fluorescence change corresponded to the rate of ATP hydrolysis, whereas tight ATP binding occurred at a much faster rate in approximate agreement with the rate of dissociation. Thus the fluorescence change in the acto-SF-1 pathway corresponds to step 3 in the SF-1 mechanism. The complete scheme can be described as follows: (formula: see text) where AM represents acto-SF-1. The tight binding step in the SF-1 pathway (k2) is sufficiently fast so that a similar step (k2') in the acto-SF-1 pathway could precede dissociation but the AM-ATP intermediate has not been detected. Following hydrolysis on the free SF-1, actin recombines with M.ADP.Pi or possibly with a second SF-1 product intermediate as proposed by Chock et al. (1976) and the fluorescence returns to the original AM level with product release.  相似文献   

11.
The elementary steps of contraction in rabbit fast twitch muscle fibers were investigated with particular emphasis on the mechanism of phosphate (Pi) binding/release, the mechanism of force generation, and the relation between them. We monitor the rate constant 2 pi b of a macroscopic exponential process (B) by imposing sinusoidal length oscillations. We find that the plot of 2 pi b vs. Pi concentration is curved. From this observation we infer that Pi released is a two step phenomenon: an isomerization followed by the actual Pi release. Our results fit well to the kinetic scheme: [formula: see text] where A = actin, M = myosin, S = MgATP (substrate), D = MgADP, P = phosphate, and Det is a composite of all the detached and weakly attached states. For our data to be consistent with this scheme, it is also necessary that step 4 (isomerization) is observed in process (B). By fitting this scheme to our data, we obtained the following kinetic constants: k4 = 56 s-1, k-4 = 129 s-1, and K5 = 0.069 mM-1, assuming that K2 = 4.9. Experiments were performed at pCa 4.82, pH 7.00, MgATP 5 mM, free ATP 5 mM, ionic strength 200 mM in K propionate medium, and at 20 degrees C. Based on these kinetic constants, we calculated the probability of each cross-bridge state as a function of Pi, and correlated this with the isometric tension. Our results indicate that all attached cross-bridges support equal amount of tension. From this, we infer that the force is generated at step 4. Detailed balance indicates that 50-65% of the free energy available from ATP hydrolysis is transformed to work at this step. For our data to be consistent with the above scheme, step 6 must be the slowest step of the cross-bridge cycle (the rate limiting step). Further, AM*D is a distinctly different state from the AMD state that is formed by adding D to the bathing solution. From our earlier ATP hydrolysis data, we estimated k6 to be 9 s-1.  相似文献   

12.
Sulfate is a partial inhibitor at low and a non-essential activator at high [ATP] of the ATPase activity of F(1). Therefore, a catalytically-competent ternary F(1) x ATP x sulfate complex can be formed. In addition, the ANS fluorescence enhancement driven by ATP hydrolysis in submitochondrial particles is also stimulated by sulfate, clearly showing that the ATP hydrolysis in its presence is coupled to H(+) translocation. However, sulfate is a strong linear inhibitor of the mitochondrial ATP synthesis. The inhibition was competitive (K (i) = 0.46 mM) with respect to Pi and mixed (K (i) = 0.60 and K'(i) = 5.6 mM) towards ADP. Since it is likely that sulfate exerts its effects by binding at the Pi binding subdomain of the catalytic site, we suggest that the catalytic site involved in the H(+) translocation driven by ATP hydrolysis has a more open conformation than the half-closed one (beta(HC)), which is an intermediate in ATP synthesis. Accordingly, ATP hydrolysis is not necessarily the exact reversal of ATP synthesis.  相似文献   

13.
We have used the technique of phosphate: water oxygen exchange to measure the rate of ATP and Pi release and Pi binding to myosin subfragment 1 and actomyosin subfragment 1 from rabbit skeletal muscle. The oxygen exchange distributions for ATP and Pi release fit a simple kinetic model with a single set of rate constants for each step. For actomyosin subfragment 1 (20 degrees C, pH 7.0, I = 50 mM), the rate constant governing ATP release is approximately 8 s-1, Pi release is at approximately 60 s-1 and Pi rebinds to an ADP state at greater than 120 M-1 s-1. These rate constants are similar to those that may occur for undistorted cross-bridges within glycerinated rabbit psoas fibers (Bowater, R., Webb, M. R., and Ferenczi, M. A. (1989) J. Biol. Chem. 264, 7193-7201.  相似文献   

14.
Kinetic evidence for multiple dynein ATPase sites   总被引:7,自引:0,他引:7  
We have examined the kinetics of ATP-induced dissociation of the microtubule-dynein complex at low ATP concentrations in the presence of vanadate, which inhibits the enzyme after the binding and hydrolysis of a single ATP per site (Shimizu, T., and Johnson, K. A. (1983) J. Biol. Chem. 258, 13833-13840). Four aspects of the dissociation reaction could not be explained by a model of dynein with a single ATP-sensitive microtubule binding site. First, titration of the light-scattering amplitude versus ATP concentration in the presence of vanadate gave Mr = 720,000/ATP binding site, indicating approximately 2.8 sites/2 million molecular weight particle. Second, the dissociation reaction was incomplete at concentrations of less than 2 microM ATP in the absence of vanadate, while the addition of vanadate led to complete dissociation at an increased rate. Third, the time course of dissociation induced by less than or equal to 1 microM ATP in the presence of vanadate was biphasic, with a small but distinct lag. Fourth, the ATP concentration dependence of the rate of dissociation in the absence of vanadate was concave upward at concentrations of ATP less than 5 microM, whereas the plot was linear in the presence of vanadate. These data suggest that dynein has three ATP-sensitive microtubule binding sites and each site must bind ATP for dynein to detach from the microtubule.  相似文献   

15.
Nucleotide-free kinesin hydrolyzes ATP with burst kinetics   总被引:1,自引:0,他引:1  
Bovine brain kinesin binds ADP tightly and contains a stoichiometric amount of ADP at its active site when isolated in the presence of free Mg2+ (Hackney, D. D. (1988) Proc. Natl. Acad. Sci. U.S.A. 85, 6314-6318). EDTA in excess of Mg2+ weakens ADP binding and nucleotide-free kinesin can be prepared by gel filtration with excess EDTA. On addition of ATP, this nucleotide-free enzyme catalyzes the rapid hydrolysis of a stoichiometric amount of ATP in a burst phase followed by much slower continued ATP hydrolysis limited by the release of ADP from the active site. This burst reaction is evident both by formation of [32P]Pi from [gamma-32P]ATP and by formation of [alpha-32P]ADP from [alpha-32P]ATP. At 1.1 nM kinesin active sites, the observed rate of the burst phase increases linearly with ATP over the 1-20 nM range yielding a bimolecular rate of net ATP binding and hydrolysis of 2.5 microM-1 s-1. The intercept at zero ATP is 0.008 s-1 which equals the ADP release rate at 0.008-0.009 s-1. This predicts a Km for ATP of approximately 3.5 nM and measurements of the dependence on ATP concentration of the steady state rate and amount of bound ADP are consistent with a Km of this magnitude.  相似文献   

16.
Energetics of the calcium-transporting ATPase   总被引:11,自引:0,他引:11  
A thermodynamic cycle for catalysis of calcium transport by the sarcoplasmic reticulum ATPase is described, based on equilibrium constants for the microscopic steps of the reaction shown in Equation 1 under a single set of experimental (formula; see text) conditions (pH 7.0, 25 degrees C, 100 mM KCl, 5 mM MgSO4): KCa = 5.9 X 10(-12) M2, K alpha ATP = 15 microM, Kint = 0.47, K alpha ADP = 0.73 mM, K'int = 1.7, K"Ca = 2.2 X 10(-6) M2, and Kp = 37 mM. The value of K"Ca was calculated by difference, from the free energy of hydrolysis of ATP. The spontaneous formation of an acylphosphate from Pi and E is made possible by the expression of 12.5 kcal mol-1 of noncovalent binding energy in E-P. Only 1.9 kcal mol-1 of binding energy is expressed in E X Pi. There is a mutual destabilization of bound phosphate and calcium in E-P X Ca2, with delta GD = 7.6 kcal mol-1, that permits transfer of phosphate to ADP and transfer of calcium to a concentrated calcium pool inside the vesicle. It is suggested that the ordered kinetic mechanism for the dissociation of E-P X Ca2, with phosphate transfer to ADP before calcium dissociation outside and phosphate transfer to water after calcium dissociation inside, preserves the Gibbs energies of these ligands and makes a major contribution to the coupling in the transport process. A lag (approximately 5 ms) before the appearance of E-P after mixing E and Pi at pH 6 is diminished by ATP and by increased [Pi]. This suggests that ATP accelerates the binding of Pi. The weak inhibition by ATP of E-P formation at equilibrium also suggests that ATP and phosphate can bind simultaneously to the enzyme at pH 6. Rate constants are greater than or equal to 115 s-1 for all the steps in the reaction sequence to form E-32P X Ca2 from E-P, Ca2+ and [32P]ATP at pH 7. E-P X Ca2 decomposes with kappa = 17 s-1, which shows that it is a kinetically competent intermediate. The value of kappa decreases to 4 s-1 if the intermediate is formed in the presence of 2 mM Ca2+. This decrease and inhibition of turnover by greater than 0.1 mM Ca2+ may result from slow decomposition of E-P X Ca3.  相似文献   

17.
Activation of the dynein adenosinetriphosphatase by microtubules   总被引:6,自引:0,他引:6  
Previous work has indicated that following the rapid adenosine 5'-triphosphate (ATP) induced dissociation of the microtubule-dynein complex, the rate-limiting step in the ATPase cycle is product release [Johnson, K. A. (1983) J. Biol. Chem. 258, 13825-13832], which occurs at a rate of approximately 2-6 s-1. In this report we complete the analysis of the ATPase cycle by examining the effect of microtubules on the rate of product release. For these studies we used repolymerized Tetrahymena axonemal microtubules and microtubule-associated protein (MAP) free bovine brain microtubules which were shown to be free of any measureable ATPase activity. Tetrahymena 22S dynein bound to these microtubules predominantly by the ATP-sensitive site and at a rate giving an apparent second-order rate constant of (0.2-1) X 10(6) M-1 s-1, which is 50-fold greater than the rate observed with brain microtubules containing MAPs. ATP induced the rapid dissociation of the microtubule-dynein complex with an apparent second-order rate constant vs. ATP concentration equal to 1.6 X 10(6) M-1 s-1; this value is only slightly lower than that observed in the presence of MAPs. After the ATP-induced dissociation, the dynein reassociated with the microtubules following a lag period due to the time required to hydrolyze the ATP. The duration of the lag time for reassociation decreased with increasing microtubule concentration, suggesting that microtubules increased the rate of ATP turnover. Direct measurements at steady state showed that the specific activity of the dynein increased with increasing microtubule concentration.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
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   

19.
The oxygen exchange parameters for the hydrolysis of ATP by the F1-ATPase have been determined over a 140,000-fold range of ATP concentrations and a 5,000-fold range of reaction velocity. The average number of water oxygens incorporated into each Pi product ranges from a limit of about 1.02 at saturating ATP concentrations to a limit of about 3.97 at very low ATP concentrations. The latter value represents 400 reversals of hydrolysis of bound ATP prior to Pi dissociation. In accord with the binding change mechanism, this means that ATP binding at one catalytic site increases the off constant of Pi and ADP from another catalytic site by at least 20,000-fold, equivalent to the use of 6 kcal mol-1 of ATP binding energy to promote product release. The estimated rate of reversal of hydrolysis of F1-ATPase-bound ATP to bound ADP + Pi varies only about 5-fold with ATP concentration. The rate is similar that observed previously for reversal of bound ATP hydrolysis or synthesis with the membrane-bound enzyme and is greater than the rate of net ATP formation during oxidative phosphorylation. This adds to evidence that energy input or membrane components are not required for bound ATP synthesis.  相似文献   

20.
P Fromme  P Gr?ber 《FEBS letters》1990,269(1):247-251
ATP-hydrolysis was measured with thylakoid membranes during continuous illumination. The concentrations of free and enzyme-bound ATP, ADP and Pi were measured using either cold ATP, [gamma-32P]ATP or [14C]ATP. The concentration of free ATP was constant, free ADP and enzyme-bound ATP were below the detection limit. Nevertheless, [gamma-32P]ATP was bound, hydrolyzed and 32Pi was released. The ADP was not released from the enzyme but cold Pi was bound from the medium, cold ATP was resynthesized and released. A quantitative analysis gave the following rate constants: ATP-binding kATP = 2 . 10(5) M-1 s-1, ADP-release: kADP less than 10(-2)s-1, Pi-release: kPi = 0.1 s-1. These rate constants are considerably smaller than under deenergized conditions. The rate constant for the release of ATP can be estimated to be at least 0.2 s-1 under energized conditions. Obviously, energization of the membrane, i.e. protonation of the enzyme leads mainly to a decrease of the rate of ATP-binding, to an increase of the rate of ATP release and to a decrease of the rate of ADP-release.  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