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.
The thermal unfolding of myosin subfragment 1 (S1) cleaved by trypsin was studied by differential scanning calorimetry. In the absence of nucleotides, trypsin splits the S1 heavy chain into three fragments (25, 50, and 20 kDa). This cleavage has no appreciable influence on the thermal unfolding of S1 examined in the presence of ADP, in the ternary complexes of S1 with ADP and phosphate analogs, such as orthovanadate (Vi) or beryllium fluoride (BeFx), and in the presence of F-actin. In the presence of ATP and in the complexes S1.ADP.Vi or S1.ADP.BeFx, trypsin produces two additional cleavages in the S1 heavy chain: a faster cleavage in the N-terminal region between Arg23 and Ile24, and a slower cleavage at the 50 kDa fragment. It has been shown that the N-terminal cleavage strongly decreases the thermal stability of S1 by shifting the maximum of its thermal transition by about 7 degrees C to a lower temperature, from 50 degrees C to 42.4 degrees C, whereas the cleavage at both these sites causes dramatic destabilization of the S1 molecule leading to total loss of its thermal transition. Our results show that S1 with ATP-induced N-terminal cleavage is able, like uncleaved S1, to undergo global structural changes in forming the stable ternary complexes with ADP and Pi analogs (Vi, BeFx). These changes are reflected in a pronounced increase of S1 thermal stability. However, S1 cleaved by trypsin in the N-terminal region is unable, unlike S1, to undergo structural changes induced by interaction with F-actin that are expressed in a 4-5 degrees C shift of the S1 thermal transition to higher temperature. Thus, the cleavage between Arg23 and Ile24 does not significantly affect nucleotide-induced structural changes in the S1, but it prevents structural changes that occur when S1 is bound to F-actin. The results suggest that the N-terminal region of the S1 heavy chain plays an important role in structural stabilization of the entire motor domain of the myosin head, and a long-distance communication pathway may exist between this region and the actin-binding sites.  相似文献   

2.
Decameric vanadate (V(10)) inhibits the actin-stimulated myosin ATPase activity, noncompetitively with actin or with ATP upon interaction with a high-affinity binding site (K(i) = 0.27 +/- 0.05 microM) in myosin subfragment-1 (S1). The binding of V(10) to S1 can be monitored from titration with V(10) of the fluorescence of S1 labeled at Cys-707 and Cys-697 with N-iodo-acetyl-N'-(5-sulfo-1-naphthyl)ethylenediamine (IAEDANS) or 5-(iodoacetamido) fluorescein, which showed the presence of only one V(10) binding site per monomer with a dissociation constant of 0.16-0.7 microM, indicating that S1 labeling with these dyes produced only a small distortion of the V(10) binding site. The large quenching of AEDANS-labeled S1 fluorescence produced by V(10) indicated that the V(10) binding site is close to Cys-697 and 707. Fluorescence studies demonstrated the following: (i) the binding of V(10) to S1 is not competitive either with actin or with ADP.V(1) or ADP.AlF(4); (ii) the affinity of V(10) for the complex S1/ADP.V(1) and S1/ADP.AlF(4) is 2- and 3-fold lower than for S1; and (iii) it is competitive with the S1 "back door" ligand P(1)P(5)-diadenosine pentaphosphate. A local conformational change in S1 upon binding of V(10) is supported by (i) a decrease of the efficiency of fluorescence energy transfer between eosin-labeled F-actin and fluorescein-labeled S1, and (ii) slower reassociation between S1 and F-actin after ATP hydrolysis. The results are consistent with binding of V(10) to the Walker A motif of ABC ATPases, which in S1 corresponds to conserved regions of the P-loop which form part of the phosphate tube.  相似文献   

3.
Stable inactive 2 : 1 complexes of the Klebsiella pneumoniae nitrogenase components (Kp2/Kp1) were prepared with ADP or the fluorescent ADP analogue, 2'(3')-O-[N-methylanthraniloyl] ADP and AlF(4)(-) or BeF(3)(-) ions. By analogy with published crystallographic data [Schindelin et al. (1997) Nature 387, 370-376)], we suggest that the metal fluoride ions replaced phosphate at the two ATP-binding sites of the iron protein, Kp2. The beryllium (BeF(x)) and aluminium (AlF(4)(-)) containing complexes are proposed to correspond to the ATP-bound state and the hydrolytic transition states, respectively, by analogy with the equivalent complexes of myosin [Fisher et al. (1995) Biochemistry 34, 8960-8972]. (31)P NMR spectroscopy showed that during the initial stages of complex formation, MgADP bound to the complexed Kp2 in a manner similar to that reported for isolated Kp2. This process was followed by a second step that caused broadening of the (31)P NMR signals and, in the case of the AlF4- complex, slow hydrolysis of some of the excess ADP to AMP and inorganic phosphate. The purified BeFx complex contained 3.8 +/- 0.1 MgADP per mol Kp1. With the AlF(4)(-) complex, MgAMP and adenosine (from MgAMP hydrolysis) replaced part of the bound MgADP although four AlF(4)(-) ions were retained, demonstrating that full occupancy by MgADP is not required for the stability of the complex. The fluorescence emission maximum of 2'(3')-O-[N-methylanthraniloyl] ADP was blue-shifted by 6-8 nm in both metal fluoride complexes and polarization was 6-9 times that of the free analogue. The fluorescence yield of bound 2'(3')-O-[N-methylanthraniloyl] ADP was enhanced by 40% in the AlF(4)(-) complex relative to the solvent but no increase in fluorescence was observed in the BeFx complex. Resonance energy transfer from conserved tyrosine residues located in proximity to the Kp2 nucleotide-binding pocket was marked in the AlF(4)(-) complex but minimal in the BeFx fluoride complex, illustrating a clear conformational difference in the Fe protein of the two complexes. Our data indicate that complex formation during the nitrogenase catalytic cycle is a multistep process involving at least four conformational states of Kp2: similar to the free Fe protein; as initially complexed with detectable (31)P NMR; as detected in mature complexes with no detectable (31)P NMR; in the AlF(4)(-) complex in which an altered tyrosine interaction permits resonance energy transfer with 2'(3')-O-[N-methylanthraniloyl] ADP.  相似文献   

4.
The effect of F-actin upon the binding of ADP to rabbit skeletal muscle myosin, heavy meromyosin, and subfragment 1 was studied by equilibrium dialysis, ultracentrifuge transport, and light scattering techniques. Both myosin and H-meromyosin (HMM) bind a maximum of approximately 1.6 mol of ADP/mol of protein, while S-1 binds approximately 0.9 mol of ADP/mol of protein. The affinity for ADP of all three preparations was similar at a given ionic strength (approximately 10(6) M-1 at 0.05 M KCl) and decreased with increasing ionic strength. Under conditions similar to those used for the measurement of ADP binding, the binding sites of myosin, HMM, and subfragment 1 (S-1) are saturated with actin at molar ratios of 2, 2, and 1 mol of actin monomer/mol of protein, respectively, as determined by light scattering, ultracentrifuge transport, and in the case of myosin by ATPase measurements. F-actin was found to inhibit ADP binding, but even at an actin concentration at least twice that required for saturation of myosin, HMM, or S-1, significant ADP binding remained. This ADP binding was inhibited by 10(-4) M pyrophosphate. The observations are consistent with the formation of an actomyosin-ADP complex in which actin and ADP are bound to myosin at distinct but interacting sites.  相似文献   

5.
The alpha-helix containing the thiols, SH1 (Cys-707) and SH2 (Cys-697), has been proposed to be one of the structural elements responsible for the transduction of conformational changes in the myosin head (subfragment-1 (S1)). Previous studies, using a method that isolated and measured the rate of the SH1-SH2 cross-linking step, showed that this helix undergoes ligand-induced conformational changes. However, because of long incubation times required for the formation of the transition state complexes (S1.ADP.BeF(x), S1.ADP.AlF(4)-, and S1.ADP.V(i)), this method could not be used to determine the cross-linking rate constants for such states. In this study, kinetic data from the SH1-SH2 cross-linking reaction were analyzed by computational methods to extract rate constants for the two-step mechanism. For S1.ADP.BeF(x), the results obtained were similar to those for S1.ATPgammaS. For reactions involving S1.ADP.AlF(4)- and S1.ADP.V(i), the first step (SH1 modification) is rate limiting; consequently, only lower limits could be established for the rate constants of the cross-linking step. Nevertheless, these results show that the cross-linking rate constants in the transition state complexes are increased at least 20-fold for all the reagents, including the shortest one, compared with nucleotide-free S1. Thus, the SH1-SH2 helix appears to be destabilized in the post-hydrolysis state.  相似文献   

6.
A new fluorescent ribose-modified ATP analogue, 2'(3')-O-[6-(N-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)amino)hexanoic]-ATP (NBD-ATP), was synthesized and its interaction with skeletal muscle myosin subfragment-1 (S-1) was studied. NBD-ATP was hydrolysed by S-1 at a rate and with divalent cation-dependence similar to those in the case of regular ATP. Skeletal HMM supported actin translocation using NBD-ATP and the velocity was slightly higher than that in the case of regular ATP. The addition of S1 to NBD-ATP resulted in quenching of NBD fluorescence. Recovery of the fluorescence intensity was noted after complete hydrolysis of NBD-ATP to NBD-ADP. The quenching of NBD-ATP fluorescence was accompanied by enhancement of intrinsic tryptophan fluorescence. These results suggested that the quenching of NBD-ATP fluorescence reflected the formation of transient states of ATPase. The formation of S-1.NBD-ADP.BeF(n) and S-1.NBD-ADP.AlF(4)(-) complexes was monitored by following changes in NBD fluorescence. The time-course of the formation fitted an exponential profile yielding rate constants of 7.38 x 10(-2) s(-1) for BeF(n) and 1.1 x 10(-3) s(-1) for AlF(4)(-). These values were similar to those estimated from the intrinsic fluorescence enhancement of trp due to the formation of S-1.ADP.BeF(n) or AlF(4)(-) reported previously by our group. Our novel ATP analogue seems to be applicable to kinetic studies on myosin.  相似文献   

7.
Stereo electron microscopy of negatively stained images showed that myosin heads in acto-subfragment-1 (S1) covalently cross-linked with 1-ethyl-3-[3-(dimethylamino)propyl]carbodiimide were predominantly short and round when ATP was added, in contrast to their uniform tilted appearance in the rigor state. As an attempt to exclude molecules which were actually dissociated but still tethered to actin by artificial cross-links, quick-freeze deep-etch electron microscopy was coupled with the mica flake method to observe uncross-linked native acto-S1 in the presence of ATP. To maintain the low affinity S1 associated to actin in the presence of ATP, a high concentration of acto-S1 was applied to mica flakes whose absorption had been chemically modified. The image of acto-S1 with added ATP agreed well with the expected time-course of reversible dissociation and reassociation, confirming the applicability of this approach to examination of the structural changes of acto-S1. S1 molecules attached to F-actin under rigor conditions or in the presence of ADP were elongated, with the long axis tilted to F-actin. Actin-attached S1 became short and round upon addition of ATP or ADP-inorganic vanadate. Adenyl-5'-yl imidodiphosphate and inorganic pyrophosphate each partially dissociated S1 from actin, as expected.  相似文献   

8.
Myosin heads are helically ordered on the thick filament surface in relaxed muscle. In mammalian and avian filaments this helical arrangement is dependent on temperature and it has been suggested that helical order is related to ATP hydrolysis by the heads. To test this hypothesis, we have used electron microscopy and image analysis to study the ability and temperature dependence of analogs of ATP and ADP.Pi to induce helical order in tarantula thick filaments. ATP or analogs were added to rigor myofibrils or purified thick filaments at 22 degrees C and 4 degrees C and the samples negatively stained. The ADP.Pi analogs ADP.AlF4 and ADP.Vi, and the ATP analogs ADP.BeFx, AMPPNP and ATPgammaNH2, all induced helical order in tarantula thick filaments, independent of temperature. In the absence of nucleotide, or in the presence of ADP or the ATP analog, ATPgammaS, there was no helical ordering. According to crystallographic and tryptophan fluorescence studies, all of these analogs, except ATPgammaS and ADP, induce the "closed" conformation of the myosin head (in which the gamma phosphate pocket is closed). We suggest that helical order requires the closed conformation of the myosin head but is not dependent on the hydrolysis of ATP.  相似文献   

9.
In order to elucidate the molecular basis of energy transduction by myosin as a molecular motor, a fluorescent ribose-modified ATP analog 2'(3')-O-[6-(N-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)amino)hexanoyl]-ATP (NBD-ATP), was utilized to study the conformational change of the myosin motor domain during ATP hydrolysis using the fluorescence resonance energy transfer (FRET) method. The FRET efficiency from the fluorescent probe, BD- or AD-labeled at the reactive cysteine residues, SH1 (Cys 707) or SH2 (Cys697), respectively, to the NBD fluorophore in the ATP binding site was measured for several transient intermediates in the ATPase cycle. The FRET efficiency was greater than that using NBD-ADP. The FRETs for the myosin.ADP.AlF4- and myosin.ADP.BeFn ternary complexes, which mimic the M*.ADP.P(i) state and M.ATP state in the ATPase cycle, respectively, were similar to that of NBD-ATP. This suggests that both the SH1 and SH2 regions change their localized conformations to move closer to the ATPase site in the M*.ATP state and M**.ADP.P(i) state than in the M*.ADP state. Furthermore, we measured energy transfer from BD in the essential light chain to NBD in the active site. Assuming the efficiency at different states, myosin adopts a conformation such that the light chain moves closer to the active site by approximately 9 A during the hydrolysis of ATP.  相似文献   

10.
H R Trayer  I P Trayer 《Biochemistry》1988,27(15):5718-5727
Fluorescence resonance energy transfer measurements have been made between Cys-374 on actin and Cys-177 on the alkali light chain of myosin subfragment 1 (S1) using several pairs of donor-acceptor chromophores. The labeled light chain was exchanged into subfragment 1 and the resulting fluorescently labeled subfragment 1 isolated by ion-exchange chromatography on SP-Trisacryl. The efficiency of energy transfer was measured by steady-state fluorescence in a strong binding complex of acto-S1 and found to represent a spatial separation between the two probes of 5.6-6.3 nm. The same measurements were then made with weak binding acto-S1 complexes generated in two ways. First, actin was complexed with p-phenylenedimaleimide-S1, a stable analogue of S1-adenosine 5'-triphosphate (ATP), obtained by cross-linking the SH1 and SH2 heavy-chain thiols of subfragment 1 [Greene, L. E., Chalovich, J. M., & Eisenberg, E. (1986) Biochemistry 25, 704-709]. Large increases in transfer efficiency indicated that the two probes had moved closer together by some 3 nm. Second, weak binding complexes were formed between subfragment 1 and actin in the presence of the regulatory proteins troponin and tropomyosin, the absence of calcium, and the presence of ATP [Chalovich, J. M., & Eisenberg, E. (1982) J. Biol. Chem. 257, 2432-2437]. The measured efficiency of energy transfer again indicated that the distance between the two labeled sites had moved closer by about 3 nm. These data support the idea that there is a considerable difference in the structure of the acto-S1 complex between the weakly and strongly bound states.  相似文献   

11.
To probe the effect of nucleotide on the formation of ionic contacts between actin and the 567-578 residue loop of the heavy chain of rabbit skeletal muscle myosin subfragment 1 (S1), the complexes between F-actin and proteolytic derivatives of S1 were submitted to chemical cross-linking with 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide. We have shown that in the absence of nucleotide both 45 kDa and 5 kDa tryptic derivatives of the central 50 kDa heavy chain fragment of S1 can be cross-linked to actin, whereas in the presence of MgADP.AlF4, only the 5 kDa fragment is involved in cross-linking reaction. By the identification of the N-terminal sequence of the 5-kDa fragment, we have found that trypsin splits the 50 kDa heavy chain fragment between Lys-572 and Gly-573, the residues located within the 567-578 loop. Using S1 preparations cleaved with elastase, we could show that the residue of 567-578 loop that can be cross-linked to actin in the presence of MgADP.AlF4 is Lys-574. The observed nucleotide-dependent changes of the actin-subfragment 1 interface indicate that the 567-578 residue loop of skeletal muscle myosin participates in the communication between the nucleotide and actin binding sites.  相似文献   

12.
Bertrand R  Derancourt J  Kassab R 《Biochemistry》2000,39(47):14626-14637
We have synthesized the luminescent and fluorescent lanthanide chelate S-(2-nitro-5-thiobenzoic acid)cysteaminyldiethylenetriaminepentaacetate-5-[(2-aminoethyl)am ino ]naphthalene-1-sulfonic acid as well as the fluorescent analogue S-(2-nitro-5-thiobenzoic acid)cysteaminyl-5-carboxyfluorescein using the procedure we recently described [Bertrand, R., Capony, J.-P., Derancourt, J., and Kassab, R. (1999) Biochemistry 38, 11914-11925]. Both mixed disulfides react with the skeletal myosin motor domain (S-1) as actin site-directed agents and label exclusively and stoichiometrically Cys 540 in the hydrophobic strong actin binding helix-loop-helix motif, causing only a 1.9-2.4-fold decrease in the V(max) for acto-S-1 ATPase. The covalently attached cysteaminyl probe side chain spans maximally 17 and 8 A, respectively, and the fluorophores have different polarity, volume, and flexibility. Thus, they may provide complementary spectroscopic information on the environmental properties of this critical actin binding region. Here, we have analyzed by extrinsic fluorescence spectroscopy S-1 derivatized with the fluorescein label or with the Tb(3+) or Eu(3+) chelate of the other label to assess the conformational transitions precisely occurring at this site upon interaction with F-actin, nucleotides, or phosphate analogues. For either label, specific spectral changes of significant amplitude were obtained, identifying at least two major structural states. One was mediated by rigor binding of F-actin in the absence or presence of MgADP. It was abolished by MgATP, and it was not produced by the binding of nonpolymerizable G-actin. A modeling of the corresponding changes in the intensity and lambda(max) of the fluorescence emission spectra, achieved using the fluorescent adducts of 2-mercaptoethanol in varying concentrations of dimethylformamide, illustrates the predicted apolar nature of the strong acto-S-1 interface. A second state was promoted by the binding of ATP, AMP-PNP, ADP.AlF4, ADP. BeFx, or PP(i). It should be prevalent in the weak acto-S-1 binding complexes. The accompanying fluorescence intensity reduction, observed with each label, in both the absence and presence of F-actin, would result from a specific modification by these ligands of the probe orientation and/or solvent accessibility as suggested by acrylamide quenching experiments. It could represent the spectral manifestation of the predicted allosteric linkage from the ATPase site to the strong actin binding site of S-1 that modulates the acto-S-1 affinity. Our study offers the basis necessary for further detailed spectroscopic investigations on the conformational dynamics in solution of the stereospecific and hydrophobic actin binding motif during the skeletal cross-bridge cycle.  相似文献   

13.
S Highsmith 《Biochemistry》1989,28(16):6745-6750
Rabbit skeletal muscle myosin and myosin subfragment 1 (S1) MgATPase activities were increased 2-3-fold by the addition of a variety of molecules that contained single straight saturated 12-16-carbon chains. The nonionic detergent dodecyl nonaoxyethylene ether (C12E9) increased the activity of S1 to 50% of maximum at a free C12E9 concentration of 27 +/- 9 microM. The activation was reversible and was not due to chemical modification of S1 amino acid side chains. The Vmax for actin-activated S1 MgATPase activity was increased 3-fold by C12E9. The apparent association constant for S1 binding to pure F-actin was reduced 3-fold by C12E9. The [C12E9] dependencies of the increase in S1 and acto-S1 MgATPase activities and of the decrease in acto-S1 binding were equal, within experimental uncertainty, suggesting that a single detergent-induced S1 conformational change is sufficient to explain the results. The stoichiometry of C12E9 bound to S1 in the S1-C12E9 complex was estimated, by the S1 concentration dependence of the C12E9 activation midpoint and by the light-scattering increase when S1 and detergent were mixed, to be 7 and 57 C12E9 molecules per S1, respectively. The results are discussed in relation to possible structural aspects of the mechanism of action for S1 and acto-S1 MgATPase activities.  相似文献   

14.
Hybrid contractile apparatus was reconstituted in skeletal muscle ghost fibers by incorporation of skeletal muscle myosin subfragment 1 (S1), smooth muscle tropomyosin and caldesmon. The spatial orientation of FITC-phalloidin-labeled actin and IAEDANS-labeled S1 during sequential steps of the acto-S1 ATPase cycle was studied by measurement of polarized fluorescence in the absence or presence of nucleotides conditioning the binding affinity of both proteins. In the fibers devoid of caldesmon addition of nucleotides evoked unidirectional synchronous changes in the orientation of the fluorescent probes attached to F-actin or S1. The results support the suggestion on the multistep rotation of the cross-bridge (myosin head and actin monomers) during the ATPase cycle. The maximal cross-bridge rotation by 7 degrees relative to the fiber axis and the increase in its rigidity by 30% were observed at transition between A**.M**.ADP.Pi (weak binding) and A--.M--.ADP (strong binding) states. When caldesmon was present in the fibers (OFF-state of the thin filament) the unidirectional changes in the orientation of actin monomers and S1 were uncoupled. The tilting of the myosin head and of the actin monomer decreased by 29% and 90%, respectively. It is suggested that in the "closed" position caldesmon "freezes" the actin filament structure and induces the transition of the intermediate state of actomyosin towards the weak-binding states, thereby inhibiting the ATPase activity of the actomyosin.  相似文献   

15.
Bertrand et al. [Bertrand, R., Derancourt, J. & Kassab, R. (1995) Biochemistry 34, 9500-9507] reported that 6-[fluoresceine-5(and 6)-carboxamido] hexanoic acid succinimidyl ester (FHS) selectively modifies Lys553, which is part of the strong actin-binding site of myosin subfragment 1 (S1). We found that the reaction of FHS with Lys533 is accompanied by a decrease in the fluorescence intensity of the reagent. The rate of the FHS reaction increased with increasing pH implying that the unprotonated form of the epsilon-amino group of Lys553 reacts with FHS. Addition of 0.4 M KCl reduced the rate of reaction significantly, which indicates ionic strength-dependent changes in the structure of S1. Limited trypsinolysis of S1 before the FHS reaction also decreased the rate of the reaction showing that the structural integrity of S1 is needed for the reactivity of Lys553. ATP, ADP, ADP.BeF(x), ADP.AlF(4), ADP.V(i) and pyrophosphate significantly decreased the rate of Lys553 labelling, suggesting nucleotide-induced conformational changes in the environment of Lys553. The fluorescence emission spectrum of the Lys553-bound FH moiety and the quenching of its fluorescence by nitromethane was not influenced by nucleotides, implying that the chemical reactivity but not the accessibility of Lys553 was decreased by the nucleotide-induced conformational change. In the presence of ATP when the M(**)ADP.P(i) state of the ATPase cycle is predominantly populated, the reaction rate decreased more than in the case of the S1.ADP.AlF(4)(-) and S1.ADP.V(i) complexes, which are believed to mimic the M(**)ADP.P(i) state. This indicates that the conformation of the S1-ADP.AlF(4)(-) and S1.ADP.V(i) complexes in the vicinity of Lys553 does not resemble the structure of the M(**)ADP.P(i) state. The rate of Lys553 labelling decreased strongly in the presence of actin. The nitromethane quenching of the Lys553-bound FHS was not influenced by actin, which indicates that the reduced reaction rate is not due to steric hindrance caused by the bulky protein but by actin induced conformational changes in the vicinity of Lys553.  相似文献   

16.
Myosin subfragment 1 (S1) with SH1 (Cys(707)) and SH2 (Cys(697)) groups cross-linked by p-phenylenedimaleimide (pPDM-S1) is thought to be an analog of the weakly bound states of myosin bound to actin. The structural properties of pPDM-S1 were compared in this study to those of S1.ADP.BeF(x) and S1.ADP.AlF(4)(-), i.e., the established structural analogs of the myosin weakly bound states. To distinguish between the conformational effects of SH1-SH2 cross-linking and those due to their monofunctional modification, we used S1 with the SH1 and SH2 groups labeled with N-phenylmaleimide (NPM-S1) as a control in our experiments. The state of the nucleotide pocket was probed using a hydrophobic fluorescent dye, 3-[4-(3-phenyl-2-pyrazolin-1-yl)benzene-1-sulfonylamido]phen ylboronic acid (PPBA). Differential scanning calorimetry (DSC) was used to study the thermal stability of S1. By both methods the conformational state of pPDM-S1 was different from that of unmodified S1 in the S1.ADP.BeF(x) and S1.ADP.AlF(4)(-) complexes and closer to that of nucleotide-free S1. Moreover, BeF(x) and AlF(4)(-) binding failed to induce conformational changes in pPDM-S1 similar to those observed in unmodified S1. Surprisingly, when pPDM cross-linking was performed on S1.ADP.BeF(x) complex, ADP.BeF(x) protected to some extent the nucleotide pocket of S1 from the effects of pPDM modification. NPM-S1 behaved similarly to pPDM-S1 in our experiments. Overall, this work presents new evidence that the conformational state of pPDM-S1 is different from that of the weakly bound state analogs, S1.ADP.BeF(x) and S1.ADP.AlF(4)(-). The similar structural effects of pPDM cross-linking of SH1 and SH2 groups and their monofunctional labeling with NPM are ascribed to the inhibitory effects of these modifications on the flexibility/mobility of the SH1-SH2 helix.  相似文献   

17.
Fluorescent derivatives of phalloidin are widely used to measure filamentous actin (F-actin) levels and to stabilize F-actin. We have characterized the kinetics and affinity of binding of tetramethylrhodaminyl (TRITC)-phalloidin to rabbit skeletal muscle F-actin and to F-actin in lysates of rabbit polymorphonuclear leukocytes (PMNs). We have defined conditions where TRITC-phalloidin can be used to inhibit F-actin depolymerization and to quantify F-actin without prior fixation. By equilibrium measurements, the affinity of TRITC-phalloidin binding to rabbit skeletal muscle F-actin (pyrene labeled) or to PMN lysate F-actin was 1-4 x 10(-7) M. In both cases, the stoichiometry of binding was approximately 1:1. Kinetic measurements of TRITC-phalloidin binding to PMN lysate F-actin resulted in an association rate constant of 420 +/- 120 M-1 sec-1 and a dissociation rate constant of 8.3 +/- 0.9 x 10(-5) sec-1. The affinity calculated from the kinetic measurements (2 +/- 1 x 10(-7) M) agreed well with that obtained by equilibrium measurements. The rate with which 0.6 microM TRITC-phalloidin inhibited 0.1 microM pyrenyl F-actin depolymerization (90% inhibition in 10 sec) was much faster than the rate of binding to pyrenyl F-actin (less than 1% bound in 10 sec), suggesting that phalloidin binds to filament ends more rapidly than to the rest of the filament. We show that TRITC-phalloidin can be used to measure F-actin levels in cell lysates when G-actin is also present (i.e., in cell lysates at high concentrations) if DNase I is included to prevent phalloidin-induced polymerization.  相似文献   

18.
Actin labeled at Gln-41 with dansyl ethylenediamine (DED) via transglutaminase reaction was used for monitoring the interaction of myosin subfragment 1 (S1) with the His-40-Gly-42 site in the 38-52 loop on F-actin. Proteolytic digestions of F-actin with subtilisin and trypsin, and acto-S1 ATPase measurements on heat-treated F-actin revealed that the labeling of Gln-41 had a stabilizing effect on subdomain 2 and the actin filaments. DED on Gln-41 had no effect on the values of K(m) and Vmax of the acto-S1 ATPase and the sliding velocities of actin filaments in the in vitro motility assays. This suggests either that S1 does not bind to the 40-42 site on actin or that such binding is not functionally important. The binding of monoclonal antidansyl IgG to DED-F-actin did not affect acto-S1 binding in the absence of nucleotides, indicating that the 40-42 site does not contribute much to rigor acto-S1 binding. Myosin-induced changes in subdomain 2 on actin were manifested through an increase in the fluorescence of DED-F-actin, a decrease in the accessibility of the probe to collisional quenchers, and a partial displacement of antidansyl IgG from actin by S1. It is proposed that these changes in the 38-52 loop on actin originate from S1 binding to other myosin recognition sites on actin.  相似文献   

19.
The role of the N-terminal region of myosin light chain 1 (LC1) in actomyosin interaction was investigated using an IgG monoclonal antibody (2H2) directed against the N-terminal region of LC1. We defined the binding site of 2H2 by examining its cross-reactivity with myosin light chains from a variety of species and with synthetic oligopeptides. Our findings suggest that 2H2 is directed against the N-terminal region of LC1 which includes the trimethylated alanine residue at the N-terminus. In the presence of 2H2, the rate of actomyosin superprecipitation was reduced, although the extent was not. 2H2 caused a reduction in the Vmax of both myosin and chymotryptic S1(A1) actin-activated ATPase activity, while the Km appeared to be unaltered. The Mg(2+)-ATPase activity of myosin alone was also unaffected. Binding studies revealed that 2H2 did not prevent the formation of acto-S1 complex, either in the presence or in the absence of ATP, nor did it affect the ability of ATP to dissociate S1 from F-actin. Our findings suggest that the N-terminal region of LC1 is not essential for actin binding but is involved in modulating actin-activated ATPase activity of myosin.  相似文献   

20.
We have synthesized the mixed disulfide, S-(2-nitro-5-thiobenzoic acid) cysteaminyl-EDTA, using a rapid procedure and water-soluble chemistry. Its disulfide-thiol exchange reaction with rabbit myosin subfragment-1 (S-1), analyzed by spectrophotometry, ATPase assays, and peptide mapping, led to the incorporation of the cysteaminyl-EDTA group into only Cys 540 on the heavy chain and into the unique cysteine on the alkali light chains. The former thiol, residing in the strong actin binding site, reacted at a much faster rate with a concomitant 3-fold decrease in the V(max) for acto-S-1 ATPase but without change in the essential enzymatic functions of S-1. Upon chelation of Fe(3+) ions to the Cys 540-bound EDTA and incubation of the S-1 derivative-Fe complex with ascorbic acid at pH 7.5, the 95 kDa heavy chain underwent a conformation-dependent, single-cut oxidative fragmentation within 5-15 A of Cys 540. Three pairs of fragments were formed which, after specific fluorescent labeling and SDS-PAGE, could be positioned along the heavy chain sequence as 68 kDa-26 kDa, 62 kDa-32 kDa, and 54 kDa-40 kDa. Densitometric measurements revealed that the yield of the 54 kDa-40 kDa pair of bands, but not that for the two other pairs, was very sensitive to S-1 binding to nucleotides or phosphate analogues as well as to F-actin. In binary complexes, all the former ligands specifically lowered the yield to 40% of S-1 alone, roughly in the following order: ADP = AMP-PNP > ATP = ADP.AlF(4) > ADP.BeF(x)() > PP(i). By contrast, rigor binding to F-actin increased the yield to 130%. In the ternary acto-S-1-ADP complex, the yield was again reduced to 80%, and it fell to 25% in acto-S-1-ADP.AlF(4), the putative transition state analogue complex of the acto-S-1 ATPase. These different quantitative changes reflect distinct ligand-induced conformations of the secondary structure element whose scission generates the 54 kDa-40 kDa species. According to the S-1 crystal structure, this element could be unambiguously assigned to the switch II helix (residues 475-507) whose N-terminus lies 14.2 A from Cys 540 and would include the ligand-responsive cleavage site. This motif is thought to be crucial for the transmission of sub-nanometer structural changes at the ATPase site to both the actin site and the lever arm domain during energy transduction. Our study illustrates this novel, actin site-specific chemical proteolysis of S-1 as a direct probe of the switch II helix conformational transitions in solution most likely associated with the skeletal cross-bridge cycle.  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