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 46 毫秒
1.
C Frieden  K Patane 《Biochemistry》1985,24(15):4192-4196
The role of adenosine 5'-triphosphate (ATP) in the Mg2+-induced conformational change of rabbit skeletal muscle G-actin has been investigated by comparing actin containing bound ADP with actin containing bound ATP. As previously described [Frieden, C. (1982) J. Biol. Chem. 257, 2882-2886], N-acetyl-N'-(5-sulfo-1-naphthyl)ethylenediamine-labeled G-actin containing ATP undergoes a time-dependent Mg2+-induced fluorescence change that reflects a conformational change in the actin. Addition of Mg2+ to labeled G-actin containing ADP gives no fluorescence change, suggesting that the conformational change does not occur. The fluorescence change can be restored on the addition of ATP. Examination of the time courses of these experiments suggests that ATP must replace ADP prior to the Mg2+-induced change. The Mg2+-induced polymerization of actin containing ADP is extraordinarily slow compared to that of actin containing ATP. The lack of the Mg2+-induced conformational change, which is an essential step in the Mg2+-induced polymerization, is probably the cause for the very slow polymerization of actin containing ADP. On the other hand, at 20 degrees C, at pH 8, and in 2 mM Mg2+, the elongation rate from the slow growing end of an actin filament, measured by using the protein brevin to block growth at the fast growing end, is only 4 times slower for actin containing ADP than for actin containing ATP.  相似文献   

2.
Xanthine oxidase increases the rate of actin polymerization. This occurs at oxidase concentrations as low as 40 nM provided the concentration of the polymerizing agent is low (0.5 mM MgCl2). In the presence of 0.1 M KCl plus 1 mM MgCl2 as the polymerizing agents, xanthine oxidase does not affect the rate of the polymerization but increases significantly the rate of the conversion of F(ATP)actin into F(ADP.Pi)actin and probably also the rate of the orthophosphate release.  相似文献   

3.
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   

4.
In the rat heart the actin-bound nucleotide contained both ATP and ADP. The ratio of bound ATP to bound ADP depended on the functional state of the heart; it was higher in hearts stopped reversibly in diastole (low Ca(2+), high Mg(2+), or high K(+)), than in stimulated (inotropic agents or pacing) hearts. Immunoblotting and gel electrophoresis showed the existence of G-actin (30% of total actin) in the cytoplasm of the heart. Pure actin was isolated from rat hearts: in G-actin the bound nucleotide readily exchanged with ATP or ADP, and in F-actin the bound nucleotide did not exchange with ATP or ADP. The free and bound nucleotides were separated in the intact heart by extraction with 75% methanol at -15 degrees C. In rat hearts perfused with (32)P-labeled orthophosphate the actin-bound nucleotide rapidly exchanged with the cytoplasmic ATP. The full exchange of the bound ATP was immediate, whereas the full exchange of the bound ADP was slower. The full exchange of the bound ATP was independent of the heartbeat frequency, whereas the full exchange of the bound ADP was frequency dependent. The data suggest that the transformation of actin monomer-ATP to actin polymer-ADP is a part of the normal contraction-relaxation cycle of the rat heart.  相似文献   

5.
In our previous study [Chalovich, J. M., Greene, L. E., & Eisenberg, E. (1983) Proc. Natl. Acad. Sci. U.S.A. 80, 4909-4913], myosin subfragment 1 that was modified by having its two reactive thiol groups cross-linked by N,N'-p-phenylenedimaleimide (pPDM) was found to resemble the myosin subfragment 1-adenosine 5'-triphosphate (S-1.ATP) complex in its interaction with actin. In the present study, we examined the effect of actin on adenosine 5'-diphosphate (ADP) trapped at the active site of pPDM.S-1. Our results indicate first that, in the presence of actin, ADP is no longer trapped at the active site but exchanges rapidly with free nucleotide. Different pPDM.S-1.nucleotide complexes were then formed by exchanging nucleotide into the active site of pPDM.S-1 in the presence of actin. The binding of pPDM.S-1.ATP or pPDM.S-1.PPi to actin is virtually identical with that of unmodified S-1 in the presence of ATP. Specifically, at mu = 18 mM, 25 degrees C, pPDM.S-1.ATP or pPDM.S-1.PPi binds to unregulated actin with the same affinity as does S-1.ATP, and this binding does not appear to be affected by troponin-tropomyosin. On the other hand, pPDM.S-1.ADP and pPDM.S-1 with no bound nucleotide both show a small, but significant, difference between their binding to actin and the binding of S-1.ATP; pPDM.S-1 and pPDM.S-1.ADP both bind about 2- to 3-fold more strongly to unregulated actin than does S-1.ATP.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Myosin VIIA was cloned from rat kidney, and the construct (M7IQ5) containing the motor domain, IQ domain, and the coiled-coil domain as well as the full-length myosin VIIA (M7full) was expressed. The M7IQ5 contained five calmodulins. Based upon native gel electrophoresis and gel filtration, it was found that M7IQ5 was single-headed, whereas M7full was two-headed, suggesting that the tail domain contributes to form the two-headed structure. M7IQ5 had Mg(2+)-ATPase activity that was markedly activated by actin with K(actin) of 33 microm and V(max) of 0.53 s(-1) head(-1). Myosin VIIA required an extremely high ATP concentration for ATPase activity, ATP-induced dissociation from actin, and in vitro actin-translocating activity. ADP markedly inhibited the actin-activated ATPase activity. ADP also significantly inhibited the ATP-induced dissociation of myosin VIIA from actin. Consistently, ADP decreased K(actin) of the actin-activated ATPase. ADP decreased the actin gliding velocity, although ADP did not stop the actin gliding even at high concentration. These results suggest that myosin VIIA has slow ATP binding or low affinity for ATP and relatively high affinity for ADP. The directionality of myosin VIIA was determined by using the polarity-marked dual fluorescence-labeled actin filaments. It was found that myosin VIIA is a plus-directed motor.  相似文献   

7.
The effect of aurovertin on the binding parameters of ADP and ATP to native F1 from beef heart mitochondria in the presence of EDTA has been explored. Three exchangeable sites per F1 were titrated by ADP and ATP in the absence or presence of aurovertin. Curvilinear Scatchard plots for the binding of both ADP and ATP were obtained in the absence of aurovertin, indicating one high affinity site (Kd for ADP = 0.6-0.8 microM; Kd for ATP = 0.3-0.5 microM) and two lower affinity sites (Kd for ADP = 8-10 microM; Kd for ATP = 7-10 microM). With a saturating concentration of aurovertin capable of filling the three beta subunits of F1, the curvilinearity of the Scatchard plots was decreased for ATP binding and abolished for ADP binding, indicating homogeneity of ADP binding sites in the F1-aurovertin complex (Kd for ADP = 2 microM). When only the high affinity aurovertin site was occupied, maximal enhancement of the fluorescence of the F1-aurovertin complex was attained with 1 mol of ADP bound per mol of F1 and maximal quenching for 1 mol of ATP bound per mol of F1. When the F1-aurovertin complex was incubated with [3H]ADP followed by [14C]ATP, full fluorescence quenching was attained when ATP had displaced the previously bound ADP. In the case of the isolated beta subunit, both ADP and ATP enhanced the fluorescence of the beta subunit-aurovertin complex. The Kd values for ADP and ATP in the presence of EDTA were 0.6 mM and 3.7 mM, respectively; MgCl2 decreased the Kd values to 0.1 mM for both ADP and ATP. It is postulated that native F1 possesses three equivalent interacting nucleotide binding sites and exists in two conformations which are in equilibrium and recognize either ATP (T conformation) or ADP (D conformation). The negative interactions between the nucleotide binding sites of F1 are strongest in the D conformation. Upon addition of aurovertin, the site-site cooperativity between the beta subunits of F1 is decreased or even abolished.  相似文献   

8.
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   

9.
The actin-bound ADP was separated from cytoplasmic nucleotides by treatment of intact arterial smooth muscle with 50% ethanol. In (32)P-labeled smooth muscle the actin-bound ADP and phosphate readily exchanged with the cytoplasmic [gamma,beta-(32)P]ATP; the specific radioactivity of actin-bound ADP was equal to that of the beta-phosphate of cytoplasmic ATP and the specific radioactivity of actin-bound phosphate was equal to that of the gamma-phosphate of cytoplasmic ATP. In contrast, the exchange of the actin-bound ADP in skeletal muscle was very slow. The presence of cytoplasmic ATP was required for the exchange of the actin-bound ADP and phosphate; if ATP synthesis was inhibited the exchange was also inhibited. The extent of exchange was reduced in muscles contracted by histamine or K(+), as compared with resting muscles. The exchange was also shown in other mammalian smooth muscles, uterus, urinary bladder, and stomach. The data indicate a dynamic state of actin in smooth muscle. The data also suggest that polymerization-depolymerization of actin is part of the contraction-relaxation cycle of smooth muscle.  相似文献   

10.
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   

11.
A rotary molecular motor that can work at near 100% efficiency   总被引:6,自引:0,他引:6  
A single molecule of F1-ATPase is by itself a rotary motor in which a central gamma-subunit rotates against a surrounding cylinder made of alpha3beta3-subunits. Driven by the three betas that sequentially hydrolyse ATP, the motor rotates in discrete 120 degree steps, as demonstrated in video images of the movement of an actin filament bound, as a marker, to the central gamma-subunit. Over a broad range of load (hydrodynamic friction against the rotating actin filament) and speed, the F1 motor produces a constant torque of ca. 40 pN nm. The work done in a 120 degree step, or the work per ATP molecule, is thus ca. 80 pN nm. In cells, the free energy of ATP hydrolysis is ca. 90 pN nm per ATP molecule, suggesting that the F1 motor can work at near 100% efficiency. We confirmed in vitro that F1 indeed does ca. 80 pN nm of work under the condition where the free energy per ATP is 90 pN nm. The high efficiency may be related to the fully reversible nature of the F1 motor: the ATP synthase, of which F1 is a part, is considered to synthesize ATP from ADP and phosphate by reverse rotation of the F1 motor. Possible mechanisms of F1 rotation are discussed.  相似文献   

12.
Soluble mitochondrial ATPase (F1) from beef heart prepared in this laboratory contained approximately 1.8 mol of ADP and 0 mol of ATP/mol of F1 which were not removed by repeated precipitation of the enzyme with ammonium sulfate solution or by gel filtration in low ionic strength buffer containing EDTA. This enzyme had full coupling activity. Treatment of the enzyme with trypsin (5 mug/mg of F1 for 3 min) reduced the "tightly bound" ADP to zero, abolished coupling activity, but had no effect on the ATPase activity, stability, or membrane-binding capability of the F1. When the trypsin concentration was varied between 0 and 5 mug/mg of F1, tightly bound ADP was removed to varying degrees, and a correlation was seen between amount of residual tightly bound ADP and residual coupling activity. Gel filtration of the native F1 in high ionic strength buffer containing EDTA also caused complete loss of tightly bound ADP and coupling ability, whereas ATPase activity, stability, and membrane-binding capability were retained. The ADP-depleted F1 preparations were unable to rebind normal amounts of ADP or any ATP in simple reloading experiments. The results strongly suggest that tightly bound ADP is required for ATP synthesis and for energy-coupled ATP hydrolysis on F1. The results also suggest that ATP synthesis and energy-linked ATP hydrolysis rather than involving one nucleotide binding site on F1, involve a series or "cluster" of sites. The ATP hydrolysis site may represent one component of this cluster. The results show that nonenergy-coupled ATP hydrolysis on F1 can occur in the absence of tightly bound ADP or ATP.  相似文献   

13.
ADP-ribosylation of skeletal muscle actin by Clostridium perfringens iota toxin increased the rate of exchange of actin-bound [gamma-32P]ATP by unlabelled ATP about twofold. Increased exchange rates were observed with ATP and ATP[gamma S], much less with ADP but not with AMP or NAD. ADP-ribosylation of skeletal muscle actin reduced "basal" and Mg2+ (1 mM)-induced ATP hydrolysis by about 80%. Similar inhibition of ATP hydrolysis was observed with liver actin ADP-ribosylated by Clostridium botulinum C2 toxin. The data indicate that ADP-ribosylation of actin at Arg-177 largely affects the ATP-binding and ATPase activity.  相似文献   

14.
Filamentous muscle actin (F-actin) aggregated blood platelets while G-actin was ineffective. This aggregation could be blocked by ATP suggesting a possible role of actin-bound ADP in this process. Actin-bound ADP caused platelet aggregation at concentrations significantly lower than equivalent concentrations of free ADP. Thus, actin potentiates the aggregating action of ADP. An actin antibody or DNase I inhibited this aggregation showing the requirement of actin in this process. Like other physiological agents, Ca++ was necessary for platelet aggregation by actin. Platelets fixed in formaldehyde were not aggregated by actin showing the need for viable platelets. Since F-actin contains 1 mole of bound ADP/mole protein, it is postulated that actin potentiates ADP-induced aggregation by providing multiple interaction sites for platelets.  相似文献   

15.
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   

16.
The nucleotide state of actin (ATP, ADP-Pi, or ADP) is known to impact its interactions with other actin molecules upon polymerization as well as with multiple actin binding proteins both in the monomeric and filamentous states of actin. Recently, molecular dynamics simulations predicted that a sequence located at the interface of subdomains 1 and 3 (W-loop; residues 165–172) changes from an unstructured loop to a β-turn conformation upon ATP hydrolysis (Zheng, X., Diraviyam, K., and Sept, D. (2007) Biophys. J. 93, 1277–1283). This region participates directly in the binding to other subunits in F-actin as well as to cofilin, profilin, and WH2 domain proteins and, therefore, could contribute to the nucleotide sensitivity of these interactions. The present study demonstrates a reciprocal communication between the W-loop region and the nucleotide binding cleft on actin. Point mutagenesis of residues 167, 169, and 170 and their site-specific labeling significantly affect the nucleotide release from the cleft region, whereas the ATP/ADP switch alters the fluorescence of probes located in the W-loop. In the ADP-Pi state, the W-loop adopts a conformation similar to that in the ATP state but different from the ADP state. Binding of latrunculin A to the nucleotide cleft favors the ATP-like conformation of the W-loop, whereas ADP-ribosylation of Arg-177 forces the W-loop into a conformation distinct from those in the ADP and ATP-states. Overall, our experimental data suggest that the W-loop of actin is a nucleotide sensor, which may contribute to the nucleotide state-dependent changes in F-actin and nucleotide state-modulated interactions of both G- and F-actin with actin-binding proteins.  相似文献   

17.
The role of the bound nucleotide in the polymerization of actin.   总被引:12,自引:0,他引:12  
R Cooke 《Biochemistry》1975,14(14):3250-3256
Three mucleotides, ATP, ADP, and an unsplit-table analog of ATP (adenylyl imidodiphosphate (AMPPNP)), were bound to monomeric actin, and their effects on the rate and extent of the actin polymerization were studied. The kinetics of polymerization, assayed by the change in OD232, followed a simple exponential curve. The rates of polymerization were equal for bound ATP and AMPPNP; both of which were three to five times faster than the rate for ADP. The concentration of actin monomers in apparent equilibrium with the polymer, G(180 degrees longitude), was determined. Values of G(180 degrees longitude) in 100 mM KCl were found for different nucleotides to be: G-ATP(180 degrees longitude) = 0.7 mu-M, G-AMPPNP(180 degrees longitude) = 0.8 MU-M, and G-ADP(180 degrees longitude) = 3.4 mu-M. The equilibrium constant of the polymerization is given by K = [G(180 degrees longitude)]-minus 1 when no nucleotide is split. The polymerization of actin-ATP is more complex due to the splitting of the nucleotide and our data require that this polymerization involves more than one step. The kinetic parameters for the polymerization of actin-ATP can be explained by a simple scheme in which the nucleotide dephosphorylation occurs in a step following the polymerization step. The conclusions are: (1) the binding of ATP to actin monomer promotes polymerization slightly more than the binding of ADP, (2) actin bound ATP provides less than 4 kJ/mol of free energy to promote polymerization, and (3) the dephosphorylation of the nucleotide is not coupled to polymerization.  相似文献   

18.
利用H^+-ATP酶复合中的Fo的色氨酸荧光,观察了复合体中F1结合ATP或ADP时,Fo的荧光猝灭常数的变化结果表明F1结合ATP或ADP时Fo可得到不同的猝来常数,也就是Fo会产生不同的构象变化。这些结果说明了H^+ATP酶合ATP合成的过程中F1与Fo之间存在着构象之间的通信与传递。  相似文献   

19.
利用H+-ATP酶复合体(也称ATP合成酶)中的Fo的色氨酸荧光,观察了复合体中F1结合ATP或ADP(酶蛋白与底物分子比为1:1)时,Fo的荧光猝灭常数的变化(用竹红菌乙作为膜区蛋白荧光的猝灭剂)结果表明F1结合ATP或ADP时Fo可得到不同的猝灭常数(Ksv),也就是Fo会产生不同的构象变化。加入二价金属离子起动ATP水解反应结束后:ATP+H2O→ADP+Pi,这时可以在Fo观察到与ADP加Mg2+时相同猝灭常数Ksv;用荧光强度随时间进程变化的实验可观察到F1水解过程中导致Fo构象变化的动力学过程。这些结果说明了H+-ATP酶复合体ATP合成的过程中F1与Fo之间存在着构象之间的通信与传递。  相似文献   

20.
Chara corallina class XI myosin is by far the fastest molecular motor. To investigate the molecular mechanism of this fast movement, we performed a kinetic analysis of a recombinant motor domain of Chara myosin. We estimated the time spent in the strongly bound state with actin by measuring rate constants of ADP dissociation from actin.motor domain complex and ATP-induced dissociation of the motor domain from actin. The rate constant of ADP dissociation from acto-motor domain was >2800 s(-1), and the rate constant of ATP-induced dissociation of the motor domain from actin at physiological ATP concentration was 2200 s(-1). From these data, the time spent in the strongly bound state with actin was estimated to be <0.82 ms. This value is the shortest among known values for various myosins and yields the duty ratio of <0.3 with a V(max) value of the actin-activated ATPase activity of 390 s(-1). The addition of the long neck domain of myosin Va to the Chara motor domain largely increased the velocity of the motility without increasing the ATP hydrolysis cycle rate, consistent with the swinging lever model. In addition, this study reveals some striking kinetic features of Chara myosin that are suited for the fast movement: a dramatic acceleration of ADP release by actin (1000-fold) and extremely fast ATP binding rate.  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