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.
E Mushtaq  L E Greene 《Biochemistry》1989,28(15):6478-6482
To elucidate the structure of the cross-bridge intermediates in the actomyosin ATPase cycle, several laboratories have added both ethylene glycol and AMP-PNP to muscle fibers. These studies suggested that ethylene glycol shifts the structure of myosin.AMP-PNP toward the weak-binding conformation, i.e., toward the structure of myosin.ATP. Since only the weak-binding conformation of myosin subfragment 1 (S-1) binds with no apparent cooperativity to the troponin-tropomyosin-actin complex (regulated actin), we used this as a probe to examine the conformation of various S-1.nucleotide complexes in ethylene glycol. Our results show that ethylene glycol markedly weakens the binding strength of S-1, S-1.ADP, and S-1.AMP-PNP to actin but has almost no effect on the binding strength of S-1.ATP. As in muscle fibers, at 40% ethylene glycol, the binding strength of S-1.AMP-PNP to actin becomes very similar to the binding strength of S-1.ATP. In the presence of troponin-tropomyosin, the binding of S-1.AMP-PNP to actin shows no apparent cooperativity in 40% ethylene glycol. Therefore, our results confirm that ethylene glycol shifts the structure of the myosin.AMP-PNP toward the weak-binding conformation. However, our results also suggest that ethylene glycol has a direct effect on the regulated actin complex. This is shown by the fact that ethylene glycol markedly increases the cooperative binding of S-1.ADP to regulated actin both in the presence and in the absence of Ca2+.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Shaw MA  Ostap EM  Goldman YE 《Biochemistry》2003,42(20):6128-6135
N-Benzyl-p-toluenesulfonamide (BTS) is a small organic molecule that specifically inhibits the contraction of fast skeletal muscle fibers. To determine the mechanism of inhibition by BTS, we performed a kinetic analysis of its effects on the elementary steps of the actomyosin subfragment-1 ATPase cycle. BTS decreases the steady-state acto-S1 ATPase rate approximately 10-fold and increases the actin concentration for half-maximal activation. BTS primarily affects three of the elementary steps of the reaction pathway. It decreases the rate of P(i) release >20-fold in the absence of actin and >100-fold in the presence of actin. It decreases the rate of S1.ADP dissociation from 3.9 to 0.8 s(-)(1) while decreasing the S1.ADP dissociation constant from 2.3 to 0.8 microM. BTS weakens the apparent affinity of S1.ADP for actin, increasing the K(d) from 7.0 to 29.5 microM. ATP binding to S1, hydrolysis, and the affinity of nucleotide-free S1 for actin are unaffected by BTS. Kinetic modeling indicates that the binding of BTS to myosin depends on actin association/dissociation and on nucleotide state. Our results suggest that the reduction of the acto-S1 ATPase rate is due to the inhibition of P(i) release, and the suppression of tension is due to inhibition of P(i) release in conjunction with the decreased apparent affinity of S1.ADP.P(i) and S1.ADP for actin.  相似文献   

3.
N.C. Millar  M.A. Geeves 《FEBS letters》1983,160(1-2):141-148
The ATP-induced dissociation of actoS1 has been studied at temperatures between −10°C and +30°C in a stopped-flow apparatus using ethylene glycol as antifreeze. At temperatures at and below 0°C the observed rate of the dissociation of actin shows a hyperbolic dependence on ATP concentration. This is interpreted in terms of a rapid binding of ATP followed by an isomerisation of the ternary complex which results in actin dissociation. Ethylene glycol weakens ATP binding but the rate of the isomerisation is unaffected. The second order rate constant for the dissociation shows a break in the Arrhenius plot.  相似文献   

4.
The rates of formation and dissociation of actin-subfragment 1 and actin-heavy mero-myosin complexes were measured by using light-scatter and the change in fluorescence of N-iodoacetyl-N'-(5-sulpho-1-naphthyl)ethylenediamine (IAEDANS)-labelled acting as probes. Association rate measurements were made at low protein concentration, where the transients approximated to single exponentials with rate constants proportional to the concentration of reactant in excess. Dissociation rate measurements were made by displacing IAEDANS-actin from myosin with excess native actin and by a salt jump. The second-order rate constant of association for actin-subfragment 1 was 3 x 10(6) M-1 . s-1 in 60 mM-KCl at 13 degree C. It was decreased 10-fold in 500 mM-KCl and in 50% (v/v) glycol. It was decreased 6-fold when MgADP or Mg[beta gamma-imido]ATP bound to myosin. The dissociation rate constant was 0.012 s-1 in 60 mM-KCl at 13 degree C. It was increased 4-fold by 500 mM-KCl, 25-fold by 50% glycol, 8-fold by MgADP binding and 170-fold by Mg[beta gamma-imido]ATP binding. Ea for association was 70 kJ . mol-1 and for dissociation 35 kJ . mol-1. Heavy meromyosin associated at twice the rate observed for subfragment 1 and dissociated at less than one-twentieth of the rate for subfragment 1 (60 mM-KCl, 25 degree C), but when Mg[beta gamma-imido]ATP bound actin-heavy meromyosin dissociated at one-half the rate for subfragment 1. There were significant correlations between increase in the dissociation rate constant, decrease in binding constant and increase in magnitude of conformational change. The association rate constant did not correlate with any property of the actin-myosin complex.  相似文献   

5.
Striated muscle thin filaments adopt different quaternary structures, depending upon calcium binding to troponin and myosin binding to actin. Modification of actin subdomain 2 alters troponin-tropomyosin-mediated regulation, suggesting that this region of actin may contain important protein-protein interaction sites. We used yeast actin mutant D56A/E57A to examine this issue. The mutation increased the affinity of tropomyosin for actin 3-fold. The addition of Ca(2+) to mutant actin filaments containing troponin-tropomyosin produced little increase in the thin filament-myosin S1 MgATPase rate. Despite this, three-dimensional reconstruction of electron microscope images of filaments in the presence of troponin and Ca(2+) showed tropomyosin to be in a position similar to that found for muscle actin filaments, where most of the myosin binding site is exposed. Troponin-tropomyosin bound with comparable affinity to mutant and wild type actin in the absence and presence of calcium, and in the presence of myosin S1, tropomyosin bound very tightly to both types of actin. The mutation decreased actin-myosin S1 affinity 13-fold in the presence of troponin-tropomyosin and 2.6-fold in the absence of the regulatory proteins. The results suggest the importance of negatively charged actin subdomain 2 residues 56 and 57 for myosin binding to actin, for tropomyosin-actin interactions, and for regulatory conformational changes in the actin-troponin-tropomyosin complex.  相似文献   

6.
Ali LF  Cohen JM  Tobacman LS 《Biochemistry》2010,49(51):10873-10880
Tropomyosin is a ubiquitous actin-binding protein with an extended coiled-coil structure. Tropomyosin-actin interactions are weak and loosely specific, but they potently influence myosin. One such influence is inhibitory and is due to tropomyosin's statistically preferred positions on actin that sterically interfere with actin's strong attachment site for myosin. Contrastingly, tropomyosin's other influence is activating. It increases myosin's overall actin affinity ~4-fold. Stoichiometric considerations cause this activating effect to equate to an ~4(7)-fold effect of myosin on the actin affinity of tropomyosin. These positive, mutual, myosin-tropomyosin effects are absent if Saccharomyces cerevisiae tropomyosin replaces mammalian tropomyosin. To investigate these phenomena, chimeric tropomyosins were generated in which 38-residue muscle tropomyosin segments replaced a natural duplication within S. cerevisiae tropomyosin TPM1. Two such chimeric tropomyosins were sufficiently folded coiled coils to allow functional study. The two chimeras differed from TPM1 but in opposite ways. Consistent with steric interference, myosin greatly decreased the actin affinity of chimera 7, which contained muscle tropomyosin residues 228-265. On the other hand, myosin S1 increased by an order of magnitude the actin affinity of chimera 3, which contained muscle tropomyosin residues 74-111. Similarly, myosin S1-ADP binding to actin was strengthened 2-fold by substitution of chimera 3 tropomyosin for wild-type TPM1. Thus, a yeast tropomyosin was induced to mimic the activating behavior of mammalian tropomyosin by inserting a mammalian tropomyosin sequence. The data were not consistent with direct tropomyosin-myosin binding. Rather, they suggest an allosteric mechanism, in which myosin and tropomyosin share an effect on the actin filament.  相似文献   

7.
Actin dimer cross-linked along the long pitch of the F-actin helix by N-(4-azido)-2-nitrophenyl (ANP) was purified by gel filtration. Purified dimers were found to polymerize on increasing the ionic strength, although at reduced rate and extent in comparison with native actin. Purified actin dimer interacts with the actin-binding proteins (ABPs) deoxyribonuclease I (DNase I) and gelsolin segment-1 (G1) as analyzed by gel filtration and native gel electrophoresis. Complex formation of the actin dimer with these ABPs inhibits its ability to polymerize. The interaction with rabbit skeletal muscle myosin subfragment 1 (S1) was analyzed for polymerized actin dimer and dimer complexed with gelsolin segment 1 or DNase I by measurement of the actin-stimulated myosin S1-ATPase and gel filtration. The data obtained indicate binding of subfragment 1 to actin dimer, albeit with considerably lower affinity than to F-actin. Polymerized actin dimer was able to stimulate the S1-ATPase activity to about 50% of the level of native F-actin. In contrast, the actin dimer complexed to DNase I or gelsolin segment 1 or to both proteins was unable to significantly stimulate the S1-ATPase. Similarly, G1:dimer complex at 20 microM stimulated the rate of release of subfragment 1 bound nucleotide (mant-ADP) only 1.6-fold in comparison to about 9-fold by native F-actin at a concentration of 0.5 microM. Using rapid kinetic techniques, a dissociation constant of 2.4 x 10 (-6) M for subfragment 1 binding to G1:dimer was determined in comparison to 3 x 10 (-8) M for native F-actin under identical conditions. Since the rate of association of subfragment 1 to G1:dimer was considerably lower than to native F-actin, we suspect that the ATP-hydrolysis by S1 was catalyzed before its association to the dimer. These data suggest an altered, nonproductive mode for the interaction of subfragment 1 with the isolated long-pitch actin dimer.  相似文献   

8.
Relaxation of both smooth and skeletal muscles appears to be caused primarily by inhibition of the step associated with Pi release in the actomyosin ATPase cycle, rather than by a block in the binding of the myosin X ATP and myosin X ADP X Pi complexes to actin. In skeletal muscle, troponin-tropomyosin not only causes marked inhibition of Pi release, but it also markedly inhibits the binding of myosin subfragment-1 X ADP to actin, raising the possibility that the two phenomena are coupled in some way. In the present study we determined whether phosphorylation of smooth muscle heavy meromyosin (HMM) also affects both the binding of HMM X ADP to actin and the Pi release step. This was done by having phosphorylated and unphosphorylated HMM X ADP compete for sites on F-actin. At mu = 30 mM, phosphorylation increased the affinity of the HMM molecule for actin about 12-fold and at mu = 170 mM, there was less than a 3-fold increase in the affinity of HMM. If phosphorylation affects the binding of each head of HMM to the same extent, then phosphorylation caused about a 4- and 2-fold increase in the affinity of each head of HMM for actin at mu = 30 and 170 mM, respectively. In contrast, at both ionic strengths, phosphorylation caused more than 100-fold actin activation of the ATPase activity of smooth muscle HMM. Therefore, the marked activation of Pi release in the acto X HMM ATPase cycle upon phosphorylation of HMM is not accompanied by a comparable increase in the affinity of HMM X ADP for actin. We have also found that phosphorylation increases by only 4-fold the rate of Pi release from HMM alone. These results suggest that in smooth muscle, phosphorylation accelerates the step associated with the release of Pi both in the forward and the reverse direction without correspondingly affecting the binding of myosin X ADP to actin.  相似文献   

9.
The role of the interaction between actin and the secondary actin binding site of myosin (segment 565-579 of rabbit skeletal muscle myosin, referred to as loop 3 in this work) has been studied with proteolytically generated smooth and skeletal muscle myosin subfragment 1 and recombinant Dictyostelium discoideum myosin II motor domain constructs. Carbodiimide-induced cross-linking between filamentous actin and myosin loop 3 took place only with the motor domain of skeletal muscle myosin and not with those of smooth muscle or D. discoideum myosin II. Chimeric constructs of the D. discoideum myosin motor domain containing loop 3 of either human skeletal muscle or nonmuscle myosin were generated. Significant actin cross-linking to the loop 3 region was obtained only with the skeletal muscle chimera both in the rigor and in the weak binding states, i.e., in the absence and in the presence of ATP analogues. Thrombin degradation of the cross-linked products was used to confirm the cross-linking site of myosin loop 3 within the actin segment 1-28. The skeletal muscle and nonmuscle myosin chimera showed a 4-6-fold increase in their actin dissociation constant, due to a significant increase in the rate for actin dissociation (k(-)(A)) with no significant change in the rate for actin binding (k(+A)). The actin-activated ATPase activity was not affected by the substitutions in the chimeric constructs. These results suggest that actin interaction with the secondary actin binding site of myosin is specific for the loop 3 sequence of striated muscle myosin isoforms but is apparently not essential either for the formation of a high affinity actin-myosin interface or for the modulation of actomyosin ATPase activity.  相似文献   

10.
T Chen  D Applegate  E Reisler 《Biochemistry》1985,24(20):5620-5625
Chemical cross-linking of actin to the 20K and 50K fragments of tryptically cleaved myosin subfragment 1 (S-1) by the zero-length cross-linking reagent 1-ethyl-3-[3-dimethylamino)propyl]carbodiimide (EDC) was used as a probe of the acto-S-1 interface in the presence of nucleotides. The course of the two reactions was monitored by measuring on sodium dodecyl sulfate (SDS)-polyacrylamide gels the time-dependent formation of the 20K-actin and 50K-actin cross-linked products. Both reactions were inhibited somewhat in the presence of MgADP, were slowed 3-4-fold in the presence of magnesium 5'-adenylyl imidodiphosphate (MgAMPPNP), and proceeded at least 7-fold slower with N,N'-p-phenylenedimaleimide (pPDM) modified S-1, as compared to the respective rates in the absence of nucleotides. However, neither the binding of the nucleotides MgADP and MgAMPPNP to S-1 nor the modification of S-1 by pPDM significantly changed the ratio of the cross-linking rates of actin to the 20K and 50K fragments. Similar to what was previously observed in the absence of nucleotides [Chen, T., Applegate, D., & Reisler, E. (1985) Biochemistry 24, 137-144], actin was cross-linked at an approximately 3-fold faster rate to the 20K fragment than to the 50K fragment under all reaction conditions tested. Thus, irrespective of the extent of acto-S-1 dissociation or the binding of nucleotides to acto-S-1, the 20K fragment remains the preferred cross-linking site for actin. These results show that the interaction of actin with each of the cross-linking sites on S-1 is not under selective or preferential control by nucleotides.  相似文献   

11.
We have used actin labelled in Cys-374 with N-(1-pyrenyl)iodoacetamide to monitor the dynamics and equilibria of the interaction between myosin subfragment 1 and the actin-troponin-tropomyosin complex in the presence of calcium. These results are compared with those obtained for pure actin and myosin subfragment 1. The sensitivity of this fluorescent label allowed us to measure the binding affinity of myosin subfragment 1 for actin directly by fluorescence titration. The affinity of subfragment 1 for actin is increased sixfold by troponin-tropomyosin in the presence of calcium. Kinetic studies of the interaction of subfragment 1 and actin have revealed an isomerization of the actin-subfragment 1 complex from a state in which actin is weakly bound (Ka = 5.9 X 10(4) M-1) to a more tightly bound complex (Ka = 1.7 X 10(7) M-1) (Coates, Criddle & Geeves (1985) Biochem. J. 232, 351). Results in the presence of troponin-tropomyosin show the same isomerization. The sixfold increase in affinity of subfragment 1 for actin is shown to be due to a decrease in the rate of dissociation of actin from the weakly bound complex.  相似文献   

12.
Smooth and non-muscle tropomyosins were found to produce a 2-3-fold Ca-insensitive stimulation of the ATPase activity of reconstituted skeletal muscles actomyosin at normal MgATP concentrations and physiological ratios of myosin to actin. Under the same conditions skeletal muscles tropomyosin had no effect. Similar effects of these three tropomyosins were observed for the low myosin/F-actin ratios necessary for kinetic measurements. Since it could be established that this actomyosin system, with or without tropomyosin, obeyed Michaelian kinetics, the tropomyosin effects could be interpreted in terms of their influence on maximal turnover (V) or on the affinity of myosin for actin (Kapp). Accordingly, gizzard tropomyosin had practically no effect on the affinity and reduced only slightly the value of V, compared to pure actin. In contrast to gizzard tropomyosin, brain tropomyosin produced an approximately twofold increase in both Kapp and V; i.e. it increased the turnover rate but decreased the affinity. It is apparent from the data that brain tropomyosin acts as an uncompetitive activator with respect to pure actin, while having the same V as the actin plus gizzard tropomyosin complex. Further studies on these tropomyosins show that only skeletal and smooth muscle tropomyosin have similar functional properties with respect to troponin inhibition and the activation of the ATPase at low ATP concentrations. It is suggested that the noted increases in V by tropomyosin are caused by the acceleration of the dissociation of the myosin head from actin at the end point of the cross bridge movement.  相似文献   

13.
We have studied the inhibition of the contraction of glycerinated rabbit psoas muscle caused by ligands that bind to the ATPase site of myosin. Two ligands, adenosine 5' (beta, gamma-imido) triphosphate (AMPPNP) and pyrophosphate (PPi), decreased the force and stiffness developed in isometric contractions and the velocity of shortening of isotonic contractions. The force exerted by isometric fibers was measured as a function of MgATP in the presence and absence of a constant concentration of the ligands. As the MgATP concentration decreased, the inhibition of tension caused by the ligand increased, reaching approximately 50% at 25 microM MgATP and either 2 mM MgPPi or 2 mM MgAMPPNP. The maximum velocity of shortening was also measured as a function of MgATP concentration in the presence of 1 and 2 mM MgPPi and 2.5 and 5 mM MgAMPPNP. Both ligands acted as pure competitive inhibitors with Ki = 3.0 mM for PPi and 5.1 mM for MgAMPPNP. These data show that both ligands are weak inhibitors of the contraction of fibers. The results provided information on the energetics of actin-myosin-ligand states that occur in the portion of the cross-bridge cycle where MgATP binds to myosin. A simple analysis of the inhibition of velocity suggests that MgAMPPNP binds to the actomyosin complex at this step of the cycle with an effective affinity constant of approximately 2 X 10(2) M-1.  相似文献   

14.
The binding of Mg2+ X adenyl-5'-yl imidodiphosphate (Mg2+ X AMP-PNP) to rabbit skeletal myofibrils has been measured in aqueous solution and in 50% ethylene glycol in the presence and absence of Ca2+. In water, the observed binding was weak with less than half the calculated myosin active sites filled even at 1 mM Mg2+ X AMP-PNP. In 50% ethylene glycol, the binding is at least 100-fold tighter and extrapolates to the expected number of binding sites. This is contrasted to the small change seen for Mg2+ X ADP binding between the same sets of conditions. This difference between Mg2+ X AMP-PNP and Mg2+ X ADP is attributed to the strong coupling of Mg2+ X AMP-PNP binding to dissociation of myosin cross-bridges. The Ca2+ sensitivity of Mg2+ X AMP-PNP binding in 50% ethylene glycol is taken as further evidence of the thermodynamic coupling of Mg2+ X AMP-PNP binding to cross-bridge dissociation. In addition, the binding of Mg2+ X AMP-PNP in 50% ethylene glycol is biphasic while Mg2+ X ADP binding under the same conditions is not. The biphasic Mg2+ X AMP-PNP binding could be caused by either the presence of two or more classes of cross-bridges or by negative cooperativity, but the presence of only a single class of Mg2+ X ADP-binding sites implies that if multiple classes of sites are involved, they do not simply differ in steric hindrance or accessibility of the binding site as a whole. The importance of using purified AMP-PNP in the study of actomyosin X AMP-PNP complexes is discussed.  相似文献   

15.
S S Margossian  S Lowey 《Biochemistry》1978,17(25):5431-5439
The effect of ionic strength, temperature, and divalent cations on the association of myosin with actin was determined in the ultracentrifuge using scanning absorption optics. The association constant (Ka) for the binding of heavy meromyosin (HmM) to F-actin was 1 X 10(7) M-1 at 20 degrees C, in 0.10 M KCl, 0.01 M imidazole (pH 7.0), 5 MM potassium phosphate, 1 mM MgCl2, and 0.3 mM ethylene glycol bis(beta-aminoethyl ether)-N,N'-tetraacetic acid. Ka was the same for HMM prepared by trypsin or chymotrypsin. The affinity of subfragment 1 (S1) for actin under the same ionic conditions was 3 X 10(6) M-1. Varying the preparative procedure for S1 had little effect on Ka. The small difference in binding energy between HMM and S1 suggests that either only one head can bind strongly to actin at a time or that free energy is lost during the sterically unfavorable attachment of the two heads to actin.  相似文献   

16.
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   

17.
Skeletal muscle myosin is an enzyme that interacts allosterically with MgATP and actin to transduce the chemical energy from ATP hydrolysis into work. By modifying myosin structure, one can change this allosteric interaction and gain insight into its mechanism. Chemical cross-linking with N,N'-p-phenylenedimaleimide (pPDM) of Cys-697 to Cys-707 of the myosin-ADP complex eliminates activity and produces a species that resembles myosin with ATP bound (Burke et al., 1976). Nucleotide-free pPDM-modified myosin subfragment 1 (S1) was prepared, and its structural and allosteric properties were investigated by comparing the nucleotide and actin interactions of S1 to those of pPDM-S1. The structural properties of the nucleotide-free pPDM-S1 are different from those of S1 in several respects. pPDM-S1 intrinsic tryptophan fluorescence intensity is reduced 28%, indicating a large increase of an internal quenching reaction (the fluorescence intensity of the related vanadate complex of S1, S1-MgADP-Vi, is reduced by a similar degree). Tryptophan fluorescence anisotropy increases from 0.168 for S1 to 0.192 for pPDM-S1, indicating that the unquenched tryptophan population in pPDM-S1 has reduced local freedom of motion. The actin affinity of pPDM-S1 is over 6,000-fold lower than that of S1, and the absolute value of the product of the net effective electric charges at the acto-S1 interface is reduced from 8.1 esu2 for S1 to 1.6 esu2 for pPDM-S1. In spite of these changes, the structural response of pPDM-S1 to nucleotide and the allosteric communication between its ATP and actin sites remain intact.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
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   

19.
High-resolution proton nuclear magnetic resonance (1H NMR) measurements were made on myosin, heavy meromyosin (HMM), myosin subfragment 1 (S1), light meromyosin (LMM), and actin. A strong signal from amino acid side chains undergoing motions too fast to be accounted for by simple rotations of groups on a rigid backbone was obtained from myosin. Comparison of myosin, HMM, S1, and LMM showed that the mobile region is located almost entirely in S1 and accounts for approximately 22% of its structure. Adenosine triphosphate (ATP) and ATP analogues had no measurable effect on the S1 spectrum. Actin, on the other hand, quenched the internal motions of S1. When S1 was titrated with actin, an association was obtained which was in agreement with other measured values. The actin effect was reversed by adding magnesium pyrophosphate (MgPPi) or adenyl-5'-yl imidophosphate (MgAMPPNP). Quantitative treatment of the broad signals from myosin and its subfragments substantiated the existence of two flexible regions in myosin. The highly mobile portion of myosin may be located in the "swivel" between S1 and the rest of myosin or in the actin binding site or in both. These possibilites are discussed, and a new possible mechanism for muscle cross bridge elasticity is proposed.  相似文献   

20.
Two Drosophila myosin II point mutations (D45 and Mhc(5)) generate Drosophila cardiac phenotypes that are similar to dilated or restrictive human cardiomyopathies. Our homology models suggest that the mutations (A261T in D45, G200D in Mhc(5)) could stabilize (D45) or destabilize (Mhc(5)) loop 1 of myosin, a region known to influence ADP release. To gain insight into the molecular mechanism that causes the cardiomyopathic phenotypes to develop, we determined whether the kinetic properties of the mutant molecules have been altered. We used myosin subfragment 1 (S1) carrying either of the two mutations (S1(A261T) and S1(G200D)) from the indirect flight muscles of Drosophila. The kinetic data show that the two point mutations have an opposite effect on the enzymatic activity of S1. S1(A261T) is less active (reduced ATPase, higher ADP affinity for S1 and actomyosin subfragment 1 (actin · S1), and reduced ATP-induced dissociation of actin · S1), whereas S1(G200D) shows increased enzymatic activity (enhanced ATPase, reduced ADP affinity for both S1 and actin · S1). The opposite changes in the myosin properties are consistent with the induced cardiac phenotypes for S1(A261T) (dilated) and S1(G200D) (restrictive). Our results provide novel insights into the molecular mechanisms that cause different cardiomyopathy phenotypes for these mutants. In addition, we report that S1(A261T) weakens the affinity of S1 · ADP for actin, whereas S1(G200D) increases it. This may account for the suppression (A261T) or enhancement (G200D) of the skeletal muscle hypercontraction phenotype induced by the troponin I held-up(2) mutation in Drosophila.  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