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.
C Tesi  F Travers  T Barman 《Biochemistry》1990,29(7):1846-1852
The initial steps of actomyosin subfragment 1 (acto-S1) ATPase (dissociation and binding of ATP) were studied at -15 degrees C with 40% ethylene glycol as antifreeze. The dissociation kinetics were followed by light scattering in a stopped-flow apparatus, and the binding of ATP was followed by the ATP chase method in a rapid-flow quench apparatus. The data from the chase experiments were fitted to E + ATP in equilibrium (K1) E.ATP----(k2) E*ATP, where E is acto-S1 or S1. The kinetics of the binding of ATP to acto-S1 were sensitive to the degree of saturation of the actin with S1. There was a sharp transition with actin nearly saturated with S1: when the S1 to actin ratio was low, the kinetics were fast (K1 greater than 300 microM, k2 greater than 40 s-1); when it was high, they were slow (K1 = 14 microM, k2 = 2 s-1). With S1 alone K1 = 12 microM and k2 = 0.07 S-1. With acto heavy meromyosin (acto-HMM) the binding kinetics were the same as with saturated acto-S1, regardless of the HMM to actin ratio. The dissociation kinetics were independent of the S1 to actin ratio. Saturation kinetics were obtained with Kd = 460 microM and kd = 75 S-1. The data for the saturated acto-S1 could be fitted to a reaction scheme, but for lack of structural information the abrupt dependence of the ATP binding kinetics upon the S1 to actin ratio is difficult to explain.(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.
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   

4.
The adenosine 5'-O-(3-thiotriphosphate) (ATP gamma S) induced dissociation of actomyosin subfragment 1 (S1) has been investigated by monitoring the light scattering changes that occur on dissociation. We have shown that ATP gamma S dissociates acto-S1 by a mechanism similar to that of ATP but at a rate 10 times slower. The maximum rate of dissociation is limited by an isomerization of the ternary actin-S1-nucleotide complex, which has a rate of 500 s-1 for ATP gamma S and an estimated rate of 5000 s-1 for ATP (20 degrees C, 0.1 M KCl, pH 7.0). The activation energy for the isomerization is the same for ATP and ATP gamma S, and both show a break in the Arrhenius plot at 5 degrees C. The reaction between acto-S1 and ATP was also followed by the fluorescence of a pyrene group covalently attached to Cys-374. We show that the fluorescence of the pyrene group reports the isomerization step and not actin dissociation. The characterization of this isomerization is discussed in relation to force-generating models of the actomyosin cross-bridge cycle.  相似文献   

5.
The mechanism of regulation of actomyosin subfragment 1 ATPase   总被引:9,自引:0,他引:9  
The mechanism of regulation of actin-subfragment 1 nucleoside triphosphatase is described in terms of the rate and equilibrium constants of a relatively simple kinetic scheme: (Formula: see text) where T, D, and Pi are nucleoside triphosphate, nucleoside diphosphate, and inorganic phosphate, respectively; Ka, Kb, and Kc are association constants; the ki are first-order rate constants; A is regulated actin (actin-tropomyosin-troponin); and M is subfragment 1. Calcium binding to regulated actin had little effect on step 2; k2 was almost unaffected, and k-2 increased, at most, 2-fold. k-1 and k3 increased 10-20-fold for ATP and 3-5-fold for 1-N6-ethenoadenosine triphosphate as substrates. Kb and Kc increased by less than 50%, whereas Ka increased 6-10-fold. The primary effect in regulation is on the rate of a conformational change which determines the rate of dissociation of ligands bound to the active site. The measurements probably underestimate the ratio of rate constants of product dissociation for active and relaxed states of actin because of heterogeneity. The kinetic evidence can be explained by a partial steric blocking mechanism or by a conformational (nonsteric) mechanism.  相似文献   

6.
The effect of Ca2+ on the interaction of bovine cardiac myosin subfragment 1 (S-1) with actin regulated by cardiac troponin-tropomyosin was evaluated. The ratios of actin to troponin and to tropomyosin were adjusted to optimize the Ca2+-dependent regulation of the steady-state actin-activated magnesium adenosinetriphosphatase (MgATPase) rate of myosin S-1. At 25 degrees C, pH 6.9, 16 mM ionic strength, the extrapolated values for maximal adenosine 5'-triphosphate (ATP) turnover rate at saturating actin, Vmax, were 6.5 s-1 in the presence of Ca2+ and 0.24 s-1 in the absence of Ca2+. In contrast to this 27-fold regulation of ATP hydrolysis, there was negligible Ca2+-dependent regulation of cardiac myosin S-1 binding to actin. In the presence of ATP, the dissociation constant of regulated actin and cardiac myosin S-1 was 32 microM in the presence of Ca2+ and 40 microM in the presence of [ethylenebis(oxyethylenenitrilo)]tetraacetic acid. These dissociation constants are indistinguishable from the concentrations of actin needed to reach half-saturation of the myosin S-1 MgATPase rates, 37 microM actin in the presence of Ca2+ and 53 microM in its absence. Although there may be Ca2+-dependent regulation of cross-bridge binding in the intact heart, the present biochemical studies suggest that cardiac regulation critically involves other parts of the cross-bridge cycle, evidenced here by almost complete Ca2+-mediated control of the myosin S-1 MgATPase rate even when the myosin S-1 is actin-bound.  相似文献   

7.
The early steps of the Mg(2+)-ATPase activity of relaxed rabbit psoas myofibrils were studied in a buffer of near-physiological ionic strength at 4 degrees C by the rapid flow quench technique. The initial ATP binding steps were studied by the ATP chase, and the cleavage and release of product steps by the Pi burst method. The data obtained were interpreted by [formula: see text] where M represents the myosin heads with or without actin interaction. This work is a continuation of our study on Ca(2+)-activated myofibrils [Houadjeto, M., Travers, F., & Barman, T. (1992) Biochemistry 31, 1564-1569]. Here the constants obtained with relaxed myofibrils were compared with those with activated myofibrils and myosin subfragment 1 (S1). We find that whereas Ca2+ increases 80X the release of products (k4), it has little effect upon the kinetics of the initial binding and cleavage steps. As with activated myofibrils and S1, the second-order binding constant for ATP (k2/K1) was about 1 microM-1 s-1 and the ATP was bound very tightly. With activated myofibrils, it was difficult to obtain an estimate for the koff for ATP(k-2) but it is much less than kcat. Here with relaxed myofibrils we estimate k-2 less than 8 x 10(-4) s-1, which is considerably smaller than kcat (0.019 s-1) and also previous estimates for this constant. The overall Kd for ATP to relaxed myofibrils is less than 8 x 10(-10) M. With S1 this Kd is about 10(-11) M.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
We have examined the kinetics of nucleotide binding to actomyosin VI by monitoring the fluorescence of pyrene-labeled actin filaments. ATP binds single-headed myosin VI following a two-step reaction mechanism with formation of a low affinity collision complex (1/K(1)' = 5.6 mm) followed by isomerization (k(+2)' = 176 s-1) to a state with weak actin affinity. The rates and affinity for ADP binding were measured by kinetic competition with ATP. This approach allows a broader range of ADP concentrations to be examined than with fluorescent nucleotide analogs, permitting the identification and characterization of transiently populated intermediates in the pathway. ADP binding to actomyosin VI, as with ATP binding, occurs via a two-step mechanism. The association rate constant for ADP binding is approximately five times greater than for ATP binding because of a higher affinity in the collision complex (1/K(5b)' = 2.2 mm) and faster isomerization rate constant (k(+5a)' = 366 s(-1)). By equilibrium titration, both heads of a myosin VI dimer bind actin strongly in rigor and with bound ADP. In the presence of ATP, conditions that favor processive stepping, myosin VI does not dwell with both heads strongly bound to actin, indicating that the second head inhibits strong binding of the lead head to actin. With both heads bound strongly, ATP binding is accelerated 2.5-fold, and ADP binding is accelerated >10-fold without affecting the rate of ADP release. We conclude that the heads of myosin VI communicate allosterically and accelerate nucleotide binding, but not dissociation, when both are bound strongly to actin.  相似文献   

9.
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   

10.
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   

11.
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   

12.
M A Geeves 《Biochemistry》1989,28(14):5864-5871
The equilibrium and dynamics of the interaction between actin, myosin subfragment 1 (S1), and ADP have been investigated by using actin which has been covalently labeled at Cys-374 with a pyrene group. The results are consistent with actin binding to S1.ADP (M.D) in a two-step reaction, A + M.D K1 equilibrium A-M.D K2 equilibrium A.M.D, in which the pyrene fluorescence only monitors the second step. In this model, K1 = 2.3 X 10(4) M-1 (k+1 = 4.6 X 10(4) M-1 s-1) and K2 = 10 (k+2 less than or equal to 4 s-1); i.e., both steps are relatively slow compared to the maximum turnover of the ATPase reaction. ADP dissociates from both M.D and A-M.D at 2 s-1 and from A.M.D at greater than or equal to 500 s-1; therefore, actin only accelerates the release of product from the A.M.D state. This model is consistent with the actomyosin ATPase model proposed by Geeves et al. [(1984) J. Muscle Res. Cell Motil. 5, 351]. The results suggest that A-M.D cannot break down at a rate greater than 4 s-1 by dissociation of ADP, by dissociation of actin, or by isomerizing to A.M.D. It is therefore unlikely to be significantly occupied in a rapidly contracting muscle, but it may have a role in a muscle contracting against a load where the ATPase rate is markedly inhibited. Under these conditions, this complex may have a role in maintaining tension with a low ATP turnover rate.  相似文献   

13.
1. The initial steps on the myosin ATPase (EC 3.6.1.3) pathway are taken to be: (formula; see text) A two-step binding for ATP is assumed, but the evidence for it is unconvincing; because of the rapidity of the process unambiguous values for K1 and K2 are not available. 2. We investigated the myosin mechanism by the chemical flow-quench technique. Reaction mixtures containing [gamma-32P]ATP plus myosin subfragment 1 were quenched in unlabelled ATP (ATP chase) or acid (Pi burst). 3. We show that the ATP-chase method can lead directly to unambiguous values for K1 and k+2. 4. The binding process was slowed down by 40% ethylene glycol. It was studied as a function of the ATP concentration. A limiting plateau resulted, showing a two-step binding for ATP, and values for K1 and k+2 were obtained. 5. K1 and k+2 are rather sensitive to the experimental conditions. Ethylene glycol and lowering of the pH decrease both constants, but an increase in KCl concentration increases them. This suggests that the binding of ATP to myosin is of an electrostatic nature. 6. The Pi-burst method can lead directly to k+3 + k-3, but under certain conditions the kinetics are governed by K1 and k+2. This uncertainty of the interpretation of Pi-burst experiments is discussed.  相似文献   

14.
The effective rate of dissociation of 1-N6-ethenoadenosine diphosphate (epsilon ADP) from the regulated actin X subfragment 1 X epsilon ADP complex of rabbit skeletal muscle is approximately 10-15 times smaller in the absence of calcium ion compared to the presence of calcium ion. The decrease in fluorescence emission with dissociation of the bound epsilon ADP fitted two exponential terms. The evidence is consistent with a kinetic scheme in which two first-order transitions precede the dissociation step: (Formula: see text) where D is epsilon ADP, A is regulated actin, M is subfragment 1, the asterisks refer to the degree of fluorescence enhancement, and AM(D) is a collision complex in equilibrium with free epsilon ADP. Both rate constants k-2 and k-1 were reduced approximately 15-fold in the absence of calcium ion. The rate constants for the dissociation of epsilon ATP, epsilon ADP X Pi, formed in the enzyme cycle, and epsilon ADP are all reduced in the absence of calcium ion; consequently, the primary effect in calcium regulation of the actin-subfragment 1 ATPase is on the rate constant of a transition (or transitions) between actomyosin-nucleoside phosphate complexes.  相似文献   

15.
Myosin V is a cellular motor protein, which transports cargos along actin filaments. It moves processively by 36-nm steps that require at least one of the two heads to be tightly bound to actin throughout the catalytic cycle. To elucidate the kinetic mechanism of processivity, we measured the rate of product release from the double-headed myosin V-HMM using a new ATP analogue, 3'-(7-diethylaminocoumarin-3-carbonylamino)-3'-deoxy-ATP (deac-aminoATP), which undergoes a 20-fold increase in fluorescence emission intensity when bound to the active site of myosin V (Forgacs, E., Cartwright, S., Kovács, M., Sakamoto, T., Sellers, J. R., Corrie, J. E. T., Webb, M. R., and White, H. D. (2006) Biochemistry 45, 13035-13045). The kinetics of ADP and deac-aminoADP dissociation from actomyosin V-HMM, following the power stroke, were determined using double-mixing stopped-flow fluorescence. These used either deac-aminoATP as the substrate with ADP or ATP chase or alternatively ATP as the substrate with either a deac-aminoADP or deac-aminoATP chase. Both sets of experiments show that the observed rate of ADP or deac-aminoADP dissociation from the trail head of actomyosin V-HMM is the same as from actomyosin V-S1. The dissociation of ADP from the lead head is decreased by up to 250-fold.  相似文献   

16.
The fluorescence emission of 1-N6-etheno-2-aza-ATP (epsilon-aza-ATP) at 410-460 nm is enhanced approximately 8-fold upon mixing substoichiometric concentrations of epsilon-aza-ATP with bovine cardiac actomyosin-S1 or myofibrils. The time course of nucleotide fluorescence measured in a front face stopped flow cell upon mixing epsilon-aza-ATP with bovine cardiac myofibrils ([Ca2+] less than 10(-7) M) is essentially the same as that with bovine cardiac actomyosin subfragment-1. In single turnover experiments, the fluorescence rapidly rises to a maximum value, then decreases with a rate constant of 0.04 s-1 at 0 degree C to a final value that is approximately twice the level of the unbound nucleotide. At concentrations of epsilon-aza-ATP greater than 40 microM the kinetics of epsilon-aza-ATP binding is clearly biphasic for both actomyosin-S1 and myofibrils. At 0 degree C, the rate of the more rapid phase is proportional to nucleotide concentration and has a second order rate constant of 1.7 X 10(5) M-1 s-1; the rate of the slower phase extrapolates to a maximum of 4-5 s-1 at high nucleotide concentration. The rate constants for dissociation of epsilon-aza-ADP from bovine cardiac actomyosin-S1 and myofibrils were measured from the decrease in epsilon-aza-ADP fluorescence enhancement observed upon displacement by ATP to be 20 and 18 s-1, respectively, at 0 degree C. These results indicate that most of the cross-bridges in cardiac myofibrils are bound to actin and that the geometric constraints imposed upon the interaction of actin and myosin by the three-dimensional structure of the myofibril do not modify the kinetics of epsilon-aza-ATP binding or epsilon-aza-ADP dissociation.  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.
A one-dimensional kinetic Ising model is developed to describe the binding of myosin subfragment 1 (SF-1) to regulated actin. The model allows for cooperative interactions between individual actin sites with bound SF-1 ligands rather than assuming that groups of actin monomer sites change their state in a cooperative fashion. With the triplet closure approximation, the model yields a set of 16 independent differential (master) equations which may be solved numerically to yield the extent of binding as a function of time. The predictions of the model are compared with experiments on the transient binding of SF-1 to regulated actin in the presence of Ca2+ and in the absence of Ca2+ with varying amounts of SF-1 prebound to the actin filament and on the equilibrium binding of SF-1 X ADP to regulated actin in the absence of Ca2+. In all cases, the calculations fit the data to within the experimental errors. In the case of SF-1 X ADP, the results suggest that a repulsive interaction exists between adjacently bound SF-1 at the ends of two neighboring seven-site actin units.  相似文献   

19.
Studies of the kinetics of association and dissociation of the formycin nucleotides FTP and FDP with CF1 were carried out using the enhancement of formycin fluorescence. The protein used, derived from lettuce chloroplasts by chloroform induced release, contains only 4 types of subunit and has a molecular weight of 280 000. In the presence of 1.25 mM MgCl2, 1 mol of ATP or FTP is bound to the latent enzyme, with Kd = 10(-7) or 2 . 10(-7), respectively. The fluorescence emission (lambdamax 340 nm) of FTP is enhanced 3-fold upon binding, and polarization of fluorescence is markedly increased. The fluorescence changes have been used to follow FTP binding, which behaves as a bimolecular process with k1 = 2.4 . 10(4) M-1 . s-1. FTP is displaced by ATP in a process apparently involving unimolecular dissociation of FTP with K-1 = 3 . 10(-3) S-1. The ratio of rates is comparable to the equilibrium constant and no additional steps have been observed. The protein has 3 sites for ADP binding. Rates of ADP binding are similar in magnitude to those for FTP. ADP and ATP sites are at least partly competitive with one another. The kinetics of nucleotide binding are strikingly altered upon activation of the protein as an ATPase. The rate of FTP binding increases to at least 10(6) M-1 . s-1. This suggests that activation involves lowering of the kinetic barriers to substrate and product binding-dissociation and has implications for the mechanism of energy transduction in photophosphorylation.  相似文献   

20.
The temperature dependence of the kinetics of the binding of ATP to myosin subfragment-1 was studied by an ATP chase technique in a rapid-flow-quench apparatus: (formula; see text) A temperature range of 30 degrees C to -15 degrees C was obtained with ethylene glycol as antifreeze. The Arrhenius plot of k2 is discontinuous with a jump at 12 degrees C. Above the jump delta H+ = 9.5 kcal/mol, below delta H+ = 28.5 kcal/mol. Few such Arrhenius plots are recorded in the literature but they are predicted from theory. Thus, we explain our results as a phase change of the subfragment 1-ATP system at 12 degrees C. This is in agreement with certain structural studies.  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