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.
The rate of ATP hydrolysis in solutions of F-actin at steady state in 50 mM KC1, 0.1 mM CaC12 was inhibited by AMP and ADP. The inhibition was competitive with ATP (Km of about 600 microM) with Ki values of 9 microM for AMP and 44 microM for ADP. ATP hydrolysis was inhibited greater than 95% by 1 mM AMP. AMP had no effect on the time course of actin polymerization, ATP hydrolysis during polymerization, or the critical actin concentration. Simultaneous measurements of G-actin/F-actin subunit exchange and nucleotide exchange showed that nucleotide exchange occurred much more rapidly than subunit exchange; during the experiment over 50% of the F-actin-bound nucleotide was replaced when less than 1% of the F-actin subunits had exchanged. When AMP was present it was incorporated into the polymer, preventing incorporation of ADP from ATP in solution. F-actin with bound Mg2+ was much less sensitive to AMP than F-actin with bound Ca2+. These data provide evidence for an ATP hydrolysis cycle associated with direct exchange of F-actin-bound ADP for ATP free in solution independent of monomer-polymer end interactions. This exchange and hydrolysis of nucleotide may be enhanced when Ca2+ is bound to the F-actin protomers.  相似文献   

2.
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   

3.
Isolation and characterization of covalently cross-linked actin dimer   总被引:5,自引:0,他引:5  
Covalently cross-linked actin dimer was isolated from rabbit skeletal muscle F-actin reacted with phenylenebismaleimide (Knight, P., and Offer, G. (1978) Biochem. J. 175, 1023-1032). The UV spectrum of the purified cross-linked actin dimer, in a nonpolymerizing buffer, was very similar to that of native F-actin and not to the spectrum of G-actin. Cross-linked actin dimer polymerized to filaments that were indistinguishable in the electron microscope from F-actin made from native G-actin and that were similar to native F-actin in their ability to activate the Mg2+-ATPase of myosin subfragment-1. The critical concentrations of polymerization of cross-linked actin dimer in 0.5 mM and 2.0 mM MgCl2, 2 to 4 microM, and 1 to 2 microM, respectively, were similar to the values for native G-actin. Cross-linked actin dimer contained 2 mol of bound nucleotide/mol of dimer. One bound nucleotide exchanged with ATP in solution with a t 1/2 of 55 min and with ADP with a t 1/2 of 5 h. The second bound nucleotide exchanged much more slowly. The more rapidly exchangeable site contained 10 to 15% bound ADP.Pi and 85 to 90% bound ATP while the second site contained much less, if any, bound ADP.Pi. Cross-linked actin dimer had an ATPase activity in 0.5 mM MgCl2 that was 7 times greater than the ATPase activity of native G-actin and that was also stimulated by cytochalasin D. These data are discussed in relation to the possible role of ATP in actin polymerization and function with the speculation that the cross-linked actin dimer may serve simultaneously as a useful model for each of the two different ends of native F-actin.  相似文献   

4.
Current theory and experiments describing actin polymerization suggest that site-specific cleavage of bound nucleotide following F-actin filament formation causes the barbed ends of microfilaments to be capped first with ATP subunits, then with ADP bound to inorganic phosphate (ADP.Pi) at steady-state. The barbed ends of depolymerizing filaments consist of ADP subunits. The decrease in stability of the barbed-end cap accompanying the transition from ADP.Pi to ADP allows nucleotide hydrolysis and subsequent loss of Pi to regulate F-actin filament dynamics. We describe a novel computational model of nucleotide capping that simulates both the spatial and temporal properties of actin polymerization. This model has been used to test the effects of high filament concentration on the behavior of the ATP hydrolysis cycle observed during polymerization. The model predicts that under conditions of high microfilament concentration an ADP cap can appear during steady-state at the barbed ends of filaments. We show that the presence of the cap can be accounted for by a kinetic model and predict the relationship between the nucleotide concentration ratio [ATP]/[ADP], the F-actin filament concentration, and the steady-state distribution of barbed-end ADP cap lengths. The possible consequences of this previously unreported phenomenon as a regulator of cytoskeletal behavior are discussed.  相似文献   

5.
The extent of actin polymerization has been studied for samples in which the bound nucleotide of the actin was ATP, ADP, or an analog of ATP that was not split (AMPPNP). The equilibrium constants for the addition of a monomer to a polymer end were determined from the concentration of monomer coexisting with the polymer. An analysis of these results concludes that the bound ATP on G-actin provides little energy to promote the polymerization of the actin. AMPPNP was incorporated into F-actin and the interaction of F-actin · AMPPNP with myosin was studied. F-actin · AMPPNP activated the ATPase of myosin to the same extent as did F-actin · ADP. However, the rate of superprecipitation was slower in the case of F-actin · AMPPNP than in the control.  相似文献   

6.
The correlation between the time courses of actin polymerization under continuous sonication and the associated ATP hydrolysis has been studied. ATP hydrolysis was not mechanistically coupled to polymerization, i.e. not necessary for polymerization, but occurred on F-actin in a subsequent monomolecular reaction. Under sonication, polymerization was complete in 10 s while hydrolysis of ATP on the polymer required 200 s. A value of 0.023 s-1 was found for the first order rate constant of ATP hydrolysis on the polymer at 25 degrees C, pH 7.8, in the presence of 0.2 mM ATP, 0.1 mM CaCl2, and 1 mM MgCl2, independent of the F-actin concentration. The conversion of ATP X F-actin to ADP X F-actin was accompanied by an increase in fluorescence of a pyrenyl probe covalently attached to actin, consistent with a 2-fold greater fluorescence for ADP X F-actin than for ATP X F-actin, with a rate constant of 0.022 s-1. In contrast, the fluorescence of F-actin labeled with 7-chloro-4-nitrobenzeno-2-oxa-1,3-diazole did not change significantly when ATP or ADP was bound. The direct consequence of the uncoupling between polymerization and ATP hydrolysis is the formation of an ATP cap at the ends of the filaments, which maintains the stability of the polymer, while most of the filament contains bound ADP. The heterogeneity of the filament with respect to ATP and ADP results in a nonlinear relationship between the rate of elongation and the concentration of G-actin with a discontinuity at the critical concentration, where the rate of growth is zero. In this respect, F-actin in ATP behaves similarly to microtubules in GTP.  相似文献   

7.
Trypsin was used as a probe of F-actin conformation. F-actin is known to be refractory to proteolysis [Jacobson, G.R. and Rosenbusch, J.P. (1976) Proc. Natl. Acad. Sci. U.S. 73, 2742-2746]. However, here it was found that F-actin could also be digested by trypsin to a 33-kDa fragment (like G-actin) when free MgADP is present in the medium. The amounts of degradation of F-actin depended on the ADP concentration; saturation occurred at about 0.5 mM. Elimination of divalent cations from the medium completely suppressed the effect of ADP on the digestion of F-actin. Other nucleotides were also examined. The effect decreased in the order ADP greater than ATP much greater than IDP greater than GDP = UDP. Adenine, adenosine, AMP, and PPi had no effect at all. epsilon-ADP had the effect, and its fluorescence was changed on the addition of F-actin. The intrinsic tryptophan fluorescence spectrum of F-actin was ADP-dependent. These results suggest the presence of a second nucleotide interacting site on actin and that ADP interaction at this site induces conformational changes in monomeric actin molecule in F-actin filaments.  相似文献   

8.
Photoaffinity labeling of the nucleotide binding site of actin   总被引:5,自引:0,他引:5  
G Hegyi  L Szilagyi  M Elzinga 《Biochemistry》1986,25(19):5793-5798
Rabbit skeletal muscle actin was photoaffinity-labeled by the nucleotide analogue 8-azidoadenosine 5'-triphosphate. In both G-actin and F-actin about 25% covalent incorporation was achieved. The labeled actins were digested with cyanogen bromide, and the labeled peptides were isolated and sequenced. In F-actin the label was bound primarily to Lys-336, while in G-actin the label was bound to Lys-336 or to Trp-356. The results indicate that the nucleotide binding site is near the phalloidin binding site of actin [Vanderkerckhove, J., Deboben, A., Nassal, M., & Wieland, T. (1985) EMBO J. 4, 2815-2818]. The binding of the azido group to Trp-356 in G-actin but not in F-actin may indicate that a change in the conformation of actin occurs in this region.  相似文献   

9.
We prepared a probe of radiolabeled, glutaraldehyde cross-linked filamentous actin (F-actin) to study binding of actin to membranes of Dictyostelium discoideum. The probe bound to membranes or detergent extracts of membranes with a high affinity and in a saturable manner. The binding could be reduced by boiling of either the actin probe or the membranes, or by addition of excess native F-actin, but not by addition of an equivalent amount of bovine serum albumin, to the assay. The probe labeled several proteins when used to overlay sodium dodecyl sulfate gels of Dictyostelium membranes. One of these labeled proteins was a 24,000-mol-wt protein (p24), which was soluble only in the presence of a high concentration of sodium deoxycholate (5%, wt/vol) at room temperature or above. The p24 was purified by selective detergent extraction and column chromatography. When tested in a novel two-phase binding assay, p24 bound both native monomeric actin (G-actin) and F-actin in a specific manner. In this assay, G-actin bound p24 with a submicromolar affinity.  相似文献   

10.
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   

11.
An unsplitable analogue of ATP (adenylyl imidodiphosphate; AMPPNP) was incorporated into F-actin [Cooke, R. (1975) Biochemistry 14, 3250-3256]. The resulting polymers (F-actin-AMPPNP) activated the ATPase activity of myosin subfragment-1 (S1) as efficiently as normal F-actin; neither the maximum velocity at infinite actin concentration (Vmax) nor the affinity of actin to S1 in the presence of ATP (1/KATPase) changed, which indicates that the terminal phosphate of the bound nucleotide at the cleft region between the two domains of the actin molecule [Kabsch, W., Mannherz, H.G., & Suck, D. (1985) EMBO J. 4, 2113-2118] is not directly involved in a myosin binding site. However, the interaction of F-actin with troponin-tropomyosin was strongly modulated by the replacement of ADP with AMPPNP. The troponin-tropomyosin complex strongly enhanced the activation of S1-ATPase activity by F-actin-AMPPNP in the presence of Ca2+, although it has no effect on the activation by normal F-actin-ADP. KATPase was enhanced about threefold by troponin-tropomyosin in the presence of Ca2+, while Vmax was not markedly changed. F-actin-AMPPNP is highly potentiated by troponin-tropomyosin even with low S1 to actin ratios and at high ATP conditions. In the absence of Ca2+, the activation by F-actin-AMPPNP was inhibited normally by troponin-tropomyosin. The results suggest that the terminal beta-phosphate of the bound nucleotide in F-actin is located in a region which is important for regulation of the interaction with myosin.  相似文献   

12.
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   

13.
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   

14.
Deoxyribonuclease I (DNase I) forms a 1:1 complex with globular actin (G-actin) and also will depolymerize filamentous actin (F-actin) to form a 1:1 complex. The effect of DNase I on the exchange of the actin nucleotide has been investigated. When DNase I is added to G-actin, the rate of nucleotide exchange is decreased from 1.16 +/- 0.25 X 10(-4) s-1 to 0.28 +/- 0.09 X 10(-4) s-1 (0 degrees C). The presence of ATP or ADP in the actin has little effect on the rate of exchange of the nucleotide for ATP. This suggests that the weaker affinity of ADP than ATP for actin is due to a slower association rate of ADP. The rate of the nucleotide exchange in the actinDNase I complex is increased by the addition of NaCl or MgCl2. When DNase I is added to F-actin, the rate of nucleotide exchange (6.2 +/- 1.6 X 10(-4) x-1, 0 degrees C) is similar to the rate of depolymerization as measured by loss of viscosity. The actinDNase I complex formed by depolymerization of F-actin exchanges nucleotide at a 4-fold faster rate than the G-actinDNase I complex in the same ionic conditions. This and other experiments suggest that DNase I binds first to F-actin before dissociating the monomer from the filament. These results are discussed in terms of possible mechanisms of action depolymerization.  相似文献   

15.
F-actin at steady state in the presence of ATP partially depolymerized to a new steady state upon mechanical fragmentation. The increase in critical concentration with the number concentration of filaments has been quantitatively studied. The data can be explained by a model in which the preferred pathway for actin association-dissociation reactions at steady state in the presence of ATP involves binding of G-actin . ATP to filaments, ATP hydrolysis, and dissociation of G-actin . ADP which is then slowly converted to G-actin . ATP. As a consequence of the slow exchange of nucleotide on G-actin, the respective amounts of G-actin . ATP and G-actin . ADP coexisting with F-actin at steady state depend on the filament number concentration. G-actin coexisting with F-actin at zero number concentration of filaments would then consist of G-actin . ATP only, while the critical concentration obtained at infinite number of filaments would be that for G-actin . ADP. Values of 0.35 and 8 microM, respectively, were found for these two extreme critical concentrations for skeletal muscle actin at 20 degrees C, pH 7.8, 0.1 mM CaCl2, 1 mM MgCl2, and 0.2 mM ATP. The same value of 8 microM was directly measured for the critical concentration of G-actin . ADP polymerized in the presence of ADP and absence of ATP, and it was unaffected by fragmentation. These results have important implications for experiments in which critical concentrations are compared under conditions that change the filament number concentrations.  相似文献   

16.
Hegyi G  Belágyi J 《The FEBS journal》2006,273(9):1896-1905
Previous cross-linking studies [Kim E, Bobkova E, Hegyi G, Muhlrad A & Reisler E (2002) Biochemistry 41, 86-93] have shown that site-specific cross-linking among F-actin monomers inhibits the motion and force generation of actomyosin. However, it does not change the steady-state ATPase parameters of actomyosin. These apparently contradictory findings have been attributed to the uncoupling of force generation from other processes of actomyosin interaction as a consequence of reduced flexibility at the interface between actin subdomains-1 and -2. In this study, we use EPR spectroscopy to investigate the effects of cross-linking constituent monomers upon the molecular dynamics of the F-actin complex. We show that cross-linking reduces the rotational mobility of an attached probe. It is consistent with the filaments becoming more rigid. Addition of heavy meromyosin (HMM) to the cross-linked filaments further restricts the rotational mobility of the probe. The effect of HMM on the actin filaments is highly cooperative: even a 1 : 10 molar ratio of HMM to actin strongly restricts the dynamics of the filaments. More interesting results are obtained when nucleotides are also added. In the presence of HMM and ADP, similar strongly reduced mobility of the probe was found than in a rigor state. In the presence of adenosine 5'[betagamma-imido] triphosphate (AMPPNP), a nonhydrolyzable analogue of ATP, weak binding of HMM to either cross-linked or native F-actin increases probe mobility. By contrast, weak binding by the HMM/ADP/AlF4 complex has different effects upon the two systems. This protein-nucleotide complex increases probe mobility in native actin filaments, as does HMM + AMPPNP. However, its addition to cross-linked filaments leaves probe mobility as constrained as in the rigor state. These findings suggest that the dynamic change upon weak binding by HMM/ADP/AlF4 which is inhibited by cross-linking is essential to the proper mechanical behaviour of the filaments during movement.  相似文献   

17.
Localization of the phalloidin and nucleotide-binding sites on actin   总被引:5,自引:0,他引:5  
Phalloidin was found to block nucleotide exchange in F-actin, without interfering with nucleotide hydrolysis. This inhibition of nucleotide exchange occurs under conditions in which monomers are able to exchange. The distance separating a fluorescent chromophore attached to phalloidin from the nucleotide on actin was determined using fluorescence resonance energy-transfer spectroscopy. They are separated by less than 1.0 nm. Added confirmation of the close proximity of phalloidin to nucleotide was obtained by extracting a small peptide-ATP complex from an actin digest. The peptide comprises residues 114-118, which are from the same region as the residues that others have shown to crosslink to phalloidin [Vandekerckhove et al. (1985) EMBO J. 4, 2815-2818]. The results suggest that phalloidin has two major effects. It traps actin monomers in a conformation which appears to be distinct from G-actin and it stabilizes the structure of F-actin, an event accompanied by the trapping of ADP.  相似文献   

18.
About 40% of the cytosolic ADP of human platelets is tightly bound to protein and the complex is precipitated from the cells by 43% ethanol. We show here that this ADP is bound to F-actin by three criteria (a) copurification with F-actin, (b) specific extraction with water and (c) by specific interaction with DNase I. Platelets contain 0.3 mumol/10(11) cells of this F-actin--ADP complex compared to the total actin content of 0.8 mumol/10(11) cells, which is consistent with the view that actin is maintained in different pools (F-actin--ADP, profilactin, G-actin). In intact platelets the F-actin-bound ADP turns over rapidly and we have determined a turnover rate at 37 degrees C of 0.1 +/- 0.025 s-1 by using a double-labelling procedure. This rapid turnover indicates that F-actin in intact platelets is in a very dynamic state, which may be necessary for rapid responses to stimuli. If it is assumed that the source of the ADP bound to F-actin is cytosolic ATP, the turnover of F-actin ADP measured represents an ATP-consuming process that would account for up to 50% of total ATP consumption in resting platelets.  相似文献   

19.
We have found that sarcoplasmic reticulum from rabbit muscle contains an ADPase which cleaves ADP bound to F-actin. The interaction is not of the simple Michaelis-Menten type, the order of the reaction being larger than the first. A possible explanation of this behaviour could be that ADPase binds to two adjacent actin monomers with a preferred orientation thus cleaving preferentially the nucleotide of one of the two monomers.  相似文献   

20.
Ikkai & Ooi [Ikkai, T. & Ooi, T. (1966) Biochemistry 5, 1551-1560] made a thorough study of the effect of pressure on G- and F-actins. However, all of the measurements in their study were made after the release of pressure. In the present experiment in situ observations were attempted by using epsilon ATP to obtain further detailed kinetic and thermodynamic information about the behaviour of actin under pressure. The dissociation rate constants of nucleotides from actin molecules (the decay curve of the intensity of fluorescence of epsilon ATP-G-actin or epsilon ADP-F-actin) followed first-order kinetics. The volume changes for the denaturation of G-actin and F-actin were estimated to be -72 mL x mol(-1) and -67 mL x mol(-1) in the presence of ATP, respectively. Changes in the intensity of fluorescence of F-actin whilst under pressure suggested that epsilon ADP-F-actin was initially depolymerized to epsilon ADP-G-actin; subsequently there was quick exchange of the epsilon ADP for free epsilon ATP, and then polymerization occurred again with the liberation of phosphate from epsilon ATP bound to G-actin in the presence of excess ATP. In the higher pressure range (> 250 MPa), the partial collapse of the three-dimensional structure of actin, which had been depolymerized under pressure, proceeded immediately after release of the nucleotide, so that it lost the ability to exchange bound ADP with external free ATP and so was denatured irreversibly. An experiment monitoring epsilon ATP fluorescence also demonstrated that, in the absence of Mg(2+)-ATP, the dissociation of actin-heavy meromyosin (HMM) complex into actin and HMM did not occur under high pressure.  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