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 546 毫秒
1.
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   

2.
Recent crystallographic studies of motor proteins showed that the structure of the motor domains of myosin and kinesin are highly conserved. Thus, these motor proteins, which are important for motility, may share a common mechanism for generating energy from ATP hydrolysis. We have previously demonstrated that, in the presence of ADP, myosin forms stable ternary complexes with new phosphate analogues of aluminum fluoride (AlF(4)(-)) and beryllium fluoride (BeF(n)), and these stable complexes mimic the transient state along the ATPase kinetic pathway [Maruta et al. (1993) J. Biol. Chem. 268, 7093-7100]. In this study, we examined the formation of kinesin.ADP.fluorometals ternary complexes and analyzed their characteristics using the fluorescent ATP analogue NBD-ATP (2'(3')-O-[6-(N-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)amino)hexanoyl]-ADP). Our results suggest that these ternary complexes may mimic transient state intermediates in the kinesin ATPase cycle. Thus, the kinesin.ADP.AlF(4)(-) complex resembles the kinesin.ADP state, and the kinesin.ADP.BeF(n) complex mimics the kinesin.ADP.P(i) state.  相似文献   

3.
Xu S  Gu J  Belknap B  White H  Yu LC 《Biophysical journal》2006,91(9):3370-3382
When myosin is attached to actin in a muscle cell, various structures in the filaments are formed. The two strongly bound states (A*M*ADP and A*M) and the weakly bound A*M*ATP states are reasonably well understood. The orientation of the strongly bound myosin heads is uniform ("stereospecific" attachment), and the attached heads exhibit little spatial fluctuation. In the prehydrolysis weakly bound A*M*ATP state, the orientations of the attached myosin heads assume a wide range of azimuthal and axial angles, indicating considerable flexibility in the myosin head. The structure of the other weakly bound state, A*M*ADP*P(i), however, is poorly understood. This state is thought to be the critical pre-power-stroke state, poised to make the transition to the strongly binding, force-generating states, and hence it is of particular interest for understanding the mechanism of contraction. However, because of the low affinity between myosin and actin in the A*M*ADP*P(i) state, the structure of this state has eluded determination both in isolated form and in muscle cells. With the knowledge recently gained in the structures of the weakly binding M*ATP, M*ADP*P(i) states and the weakly attached A*M*ATP state in muscle fibers, it is now feasible to delineate the in vivo structure of the attached state of A*M*ADP*P(i). The series of experiments presented in this article were carried out under relaxing conditions at 25 degrees C, where approximately 95% of the myosin heads in the skinned rabbit psoas muscle contain the hydrolysis products. The affinity for actin is enhanced by adding polyethylene glycol (PEG) or by lowering the ionic strength in the bathing solution. Solution kinetics and binding constants were determined in the presence and in the absence of PEG. When the binding between actin and myosin was increased, both the myosin layer lines and the actin layer lines increased in intensity, but the intensity profiles did not change. The configuration (mode) of attachment in the A*M*ADP*P(i) state is thus unique among the intermediate attached states of the cross-bridge ATP hydrolysis cycle. One of the simplest explanations is that both myosin filaments and actin filaments are stabilized (e.g., undergo reduced spatial fluctuations) by the attachment. The alignment of the myosin heads in the thick filaments and the alignment of the actin monomers in the thin filaments are improved as a result. The compact atomic structure of M*ADP*P(i) with strongly coupled domains may contribute to the unique attachment configuration: the "primed" myosin heads may function as "transient struts" when attached to the thin filaments.  相似文献   

4.
The Pgp (P-glycoprotein) multidrug transporter couples ATP hydrolysis at two cytoplasmic NBDs (nucleotide-binding domains) to the transport of hydrophobic compounds. Orthovanadate (V(i)) and fluoroaluminate (AlF(x)) trap nucleotide in one NBD by forming stable catalytically inactive complexes (Pgp-M2+-ADP-X), which are proposed to resemble the catalytic transition state, whereas the complex formed by beryllium fluoride (BeF(x)) is proposed to resemble the ground state. We studied the trapped complexes formed via incubation of Pgp with ATP (catalytically forward) or ADP (reverse) and V(i), BeF(x) or AlF(x) using Mg2+ or Co2+ as the bivalent cation. Quenching of intrinsic Pgp tryptophan fluorescence by acrylamide, iodide and caesium indicated that conformational changes took place upon formation of the trapped complexes. Trapping with V(i) and ATP led to a 6-fold increase in the acrylamide quenching constant, K(SV), suggesting that large conformational changes take place in the Pgp transmembrane regions on trapping in the forward direction. Trapping with V(i) and ADP gave only a small change in quenching, indicating that the forward- and reverse-trapped complexes are different. TNP (trinitrophenyl)-ATP/TNP-ADP interacted with all of the trapped complexes, however, the fluorescence enhancement differed for the trapped states, suggesting a change in polarity in the nucleotide-binding sites. The nucleotide-binding site of the BeF(x)-trapped complex was much more polar than that of the V(i) and AlF(x) complexes. Functionally, all the trapped complexes were able to bind drugs and TNP-nucleotides with unchanged affinity compared with native Pgp.  相似文献   

5.
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   

6.
Complexes formed from A13+ or Be2+ and fluoride inhibit the single-stranded DNA-dependent ATPase activity of RecA protein. In contrast, poly(dT)-RecA-ADP complexes, which are inactive for cleavage of LexA protein, become fully active in the presence of AlF4- or BeF3- ions. These data suggest that fluoride complexes of aluminum and beryllium (called herein X) convert RecA-ADP complexes, which bind weakly to single-stranded DNA, into RecA-ADP-X complexes, which bind tightly to single-stranded DNA, the ADP-X moiety behaving as a nonhydrolyzable analogue of ATP. We propose that AlF4- and BeF3- ions act as analogues of inorganic phosphate by binding to the site of the gamma-phosphate of ATP on RecA-ADP complexes, hence mimicking the single-stranded DNA-RecA-ADP-Pi transition state. We conclude that the elementary reaction that switches RecA protein from a high affinity single-stranded DNA binding state to a low affinity single-stranded DNA binding state is not ATP hydrolysis per se but Pi release.  相似文献   

7.
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   

8.
In order to elucidate how the elementary reactions of GTP cleavage and subsequent inorganic phosphate (Pi) release, which accompany microtubule assembly, regulate microtubule dynamics, the effect of Pi and of its structural analogues AlF4- and BeF3- on the stability of GDP-microtubules has been investigated. Inorganic phosphate binds to microtubules with a low affinity (KD = 25 mM) and slows down the rate of GDP-subunit dissociation by about 2 orders of magnitude. AlF4- and BeF3- exhibit phosphate-like effects with 1000-fold higher affinity. Evidence has been obtained for direct binding of BeF3- to microtubules with a stoichiometry of 1 mol of BeF3- per mole of GDP-subunit and an equilibrium dissociation constant of 12-15 microM. AlF4- and Pi compete for this site. Phosphate analogues abolish oscillatory polymerization kinetics and slow down microtubule turnover at steady state. In view of these results, we propose that Pi and its structural analogues bind to the site of the gamma-phosphate of GTP in the E site and reconstitute a GDP-Pi-microtubule, from which tubulin subunits dissociate very slowly. We therefore understand that, following GTP cleavage on microtubules, Pi release in the medium is accompanied by a structural change resulting in a large destabilization of the polymer. A cap of slowly dissociating GDP-Pi-subunits prevents depolymerization of the microtubule GDP-core at steady state. The similarity with the actin system [Carlier, M.-F., & Pantaloni, D. (1988) J. Biol. Chem. 263, 817-825] is underlined.  相似文献   

9.
Cooperative interaction between myosin and actin filaments has been detected by a number of different methods, and has been suggested to have some role in force generation by the actomyosin motor. In this study, we observed the binding of myosin to actin filaments directly using fluorescence microscopy to analyze the mechanism of the cooperative interaction in more detail. For this purpose, we prepared fluorescently labeled heavy meromyosin (HMM) of rabbit skeletal muscle myosin and Dictyostelium myosin II. Both types of HMMs formed fluorescent clusters along actin filaments when added at substoichiometric amounts. Quantitative analysis of the fluorescence intensity of the HMM clusters revealed that there are two distinct types of cooperative binding. The stronger form was observed along Ca2+-actin filaments with substoichiometric amounts of bound phalloidin, in which the density of HMM molecules in the clusters was comparable to full decoration. The novel, weaker form was observed along Mg2+-actin filaments with and without stoichiometric amounts of phalloidin. HMM density in the clusters of the weaker form was several-fold lower than full decoration. The weak cooperative binding required sub-micromolar ATP, and did not occur in the absence of nucleotides or in the presence of ADP and ADP-Vi. The G680V mutant of Dictyostelium HMM, which over-occupies the ADP-Pi bound state in the presence of actin filaments and ATP, also formed clusters along Mg2+-actin filaments, suggesting that the weak cooperative binding of HMM to actin filaments occurs or initiates at an intermediate state of the actomyosin-ADP-Pi complex other than that attained by adding ADP-Vi.  相似文献   

10.
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   

11.
The internal dynamics and thermal unfolding of fibre bundles prepared from rabbit psoas muscle has been studied in the presence of nucleotides by differential scanning calorimetry (DSC) and electron paramagnetic resonance (EPR) spectroscopy. Using ADP, adenosine 5'-triphosphate (ATP), AMP.PNP and inorganic phosphate analogue orthovanadate (V(i)), AlF(4)(-) and BeF(3)(-), three intermediate states of the ATP hydrolysis cycle were simulated in glycerinated muscle fibres. In the main transition of the DSC pattern, three overlapping endotherms were detected in rigor, four in strongly as well as weakly binding state of myosin to actin. Deconvolution procedure showed that the transition temperature of 67.5 degrees C was the same for rigor and strong binding state of myosin. In contrast, nucleotide binding induced shift of the melting temperatures of 52 degrees C and 67.5 degrees C, appeared a new fourth peak at 74 and 77 degrees C and produced changes in the calorimetric enthalpies. The changes of the parameters of the peak functions suggest global rearrangements of the internal structure in myosin heads in the intermediate states. In the presence of ADP or ATP plus phosphate analogue orthovanadate or beryllium fluoride, aluminium fluoride, the conventional EPR spectra of spin-labeled muscle fibres showed large changes in the ordering of the probe molecules, and a new distribution of spin labels appeared. ATP plus orthovanadate induced the orientation disorder of myosin heads; the random population of spin labels gave evidence of large local conformational and motional changes in the internal structure of myosin heads. Saturation transfer EPR measurements reported increased rotational mobility of spin labels in the presence of ATP plus phosphate analogues corresponding to weakly binding state of myosin to actin.  相似文献   

12.
We have applied differential scanning calorimetry to investigate thermal unfolding of F-actin. It has been shown that the thermal stability of F-actin strongly depends on ADP concentration. The transition temperature, T(m), increases with increasing ADP concentration up to 1 mM. The T(m) value also depends on the concentration of F-actin: it increases by almost 3 degrees C as the F-actin concentration is increased from 0.5 to 2.0 mg/ml. Similar dependence of the T(m) value on protein concentration was demonstrated for F-actin stabilized by phalloidin, whereas it was much less pronounced in the presence of AlF4(-). However, T(m) was independent of protein concentration in the case of monomeric G-actin. The results suggest that at least two reversible stages precede irreversible thermal denaturation of F-actin; one of them is dissociation of ADP from actin subunits, and another is dissociation of subunits from the ends of actin filaments. The model explains why unfolding of F-actin depends on both ADP and protein concentration.  相似文献   

13.
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   

14.
The structural natures of stable analogues for the ADP-insensitive phosphoenzyme (E2P) of Ca(2+)-ATPase formed in sarcoplasmic reticulum vesicles, i.e. the enzymes with bound beryllium fluoride (BeF.E2), bound aluminum fluoride (AlF.E2), and bound magnesium fluoride (MgF.E2), were explored and compared with those of actual E2P formed from P(i) without Ca(2+). Changes in trinitrophenyl-AMP fluorescence revealed that the catalytic site is strongly hydrophobic in BeF.E2 as in E2P but hydrophilic in MgF.E2 and AlF.E2; yet, the three cytoplasmic domains are compactly organized in these states. Thapsigargin, which was shown in the crystal structure to fix the transmembrane helices and, thus, the postulated Ca(2+) release pathway to lumen in a closed state, largely reduced the tryptophan fluorescence in BeF.E2 as in E2P, but only very slightly (hence, the release pathway is likely closed without thapsigargin) in MgF.E2 and AlF.E2 as in dephosphorylated enzyme. Consistently, the completely suppressed Ca(2+)-ATPase activity in BeF-treated vesicles was rapidly restored in the presence of ionophore A23187 but not in its absence by incubation with Ca(2+) (over several millimolar concentrations) at pH 6, and, therefore, lumenal Ca(2+) is accessible to reactivate the enzyme. In contrast, no or only very slow restoration was observed with vesicles treated with MgF and AlF even with A23187. BeF.E2 thus has the features very similar to those characteristic of the E2P ground state, although AlF.E2 and MgF.E2 most likely mimic the transition or product state for the E2P hydrolysis, during which the hydrophobic nature around the phosphorylation site is lost and the Ca(2+) release pathway is closed. The change in hydrophobic nature is probably associated with the change in phosphate geometry from the covalently bound tetrahedral ground state (BeF(3)(-)) to trigonal bipyramidal transition state (AlF(3) or AlF(4)(-)) and further to tetrahedral product state (MgF(4)(2-)), and such change likely rearranges transmembrane helices to prevent access and leakage of lumenal Ca(2+).  相似文献   

15.
The ability of adenyl-5'-yl imidodiphosphate (AMP-PNP), ADP, and PPi to dissociate the actin.myosin subfragment 1 (S-1) complex was studied using an analytical ultracentrifuge with UV optics, which enabled the direct determination of the dissociated S-1. At mu = 0.22 M, pH 7.0, 22 degrees C, with saturating nucleotide present, ADP weakens the binding of S-1 to actin about 40-fold (K congruent to 10(5) M-1), while both AMP-PNP and PPi weakens the binding about 400-fold (K congruent to 10(4) M-1). This 10-fold stronger dissociating effect of AMP-PNP and PPi compared to ADP correlates with our data showing that the binding of AMP-PNP and PPi to S-1 is about 10-fold stronger than the binding of ADP. In contrast, the binding constants of ADP, AMP-PNP, and PPi to acto.S-1 are nearly identical (K congruent to 5 x 10(3) M-1). At 4 degrees C, AMP-PNP has only a 3-fold stronger dissociating effect than ADP and, similarly, our data suggest that the binding of AMP-PNP and ADP to S-1 is quite similar at 4 degrees C. AMP-PNP and PPi are, therefore, somewhat better dissociating agents than ADP, but the difference among these three ligands is quite small. These data also show that actin and nucleotide bind to separate but interacting sites on S-1 and that the S-1 molecules bind independently along the F-actin filament with a binding constant of about 1 x 10(7) M-1 at 22 degrees C and physiological ionic strength.  相似文献   

16.
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   

17.
It is known that ternary complexes of myosin subfragment 1 (S1) with ADP and the Pi analogs beryllium fluoride (BeFx) and aluminum fluoride (AlF4-) are stable analogs of the myosin ATPase intermediates M* x ATP and M** x ADP x Pi, respectively. Using kinetic approaches, we compared the rate of formation of the complexes S1 x ADP x BeFx and S1 x ADP x AlF4- in the absence and in the presence of F-actin, as well as of the interaction of these complexes with F-actin. We show that in the absence of F-actin the formation of S1 x ADP x BeFx occurs much faster (3-4 min) than that of S1 x ADP x AlF4- (hours). The formation of these complexes in the presence of F-actin led to dissociation of S1 from F-actin, this process being monitored by a decrease in light scattering. The light scattering decrease of the acto-S1 complex occurred much faster after addition of BeFx (during 1 min) than after addition of AlF4- (more than 20 min). In both cases the light scattering of the acto-S1 complex decreased by 40-50%, but it remained much higher than that of F-actin measured in the absence of S1. The interaction of the S1 x ADP x BeFx and S1 x ADP x AlF4- complexes with F-actin was studied by the stopped-flow technique with high time resolution (no more than 0.6 sec after mixing of S1 with F-actin). We found that the binding of S1 x ADP x BeFx or S1 x ADP x AlF4- to F-actin is accompanied by a fast increase in light scattering, but it does not affect the fluorescence of a pyrene label specifically attached to F-actin. We conclude from these data that within this time range a "weak" binding of the S1 x ADP x BeFx and S1 x ADP x AlF4- complexes to F-actin occurs without the subsequent transition of the "weak" binding state to the "strong" binding state. Comparison of the light scattering kinetic curves shows that S1 x ADP x AlF4- binds to F-actin faster than S1 x ADP x BeFx does: the second-order rate constants for the "weak" binding to F-actin are (62.8 +/- 1.8) x 10(6) M-1 x sec-1 in the case of S1 x ADP x AlF4- and (22.6 +/- 0.4) x 10(6) M-1 x sec-1 in the case of S1 x ADP x BeFx. We conclude that the stable ternary complexes S1 x ADP x BeFx and S1 x ADP x AlF4- can be successfully used for kinetic studies of the "weak" binding of the myosin heads to F-actin.  相似文献   

18.
Coupling with ATP hydrolysis and cooperating with GroES, the double ring chaperonin GroEL assists the folding of other proteins. Here we report novel GroEL-GroES complexes formed in fluoroberyllate (BeF(x)) that can mimic the phosphate part of the enzyme-bound nucleotides. In ATP, BeF(x) stops the functional turnover of GroEL by preventing GroES release and produces a symmetric 1:2 GroEL-GroES complex in which both GroEL rings contain ADP.BeF(x) and an encapsulated substrate protein. In ADP, the substrate protein-loaded GroEL cannot bind GroES. In ADP plus BeF(x), however, it can bind GroES to form a stable 1:1 GroEL-GroES complex in which one of GroEL rings contains ADP.BeF(x) and an encapsulated substrate protein. This 1:1 GroEL-GroES complex is converted into the symmetric 1:2 GroEL-GroES complex when GroES is supplied in ATP plus BeF(x). Thus, BeF(x) stabilizes two GroEL-GroES complexes; one with a single folding chamber and the other with double folding chambers. These results shed light on the intermediate ADP.P(i) nucleotide states in the functional cycle of GroEL.  相似文献   

19.
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   

20.
Actin-based motility demands the spatial and temporal coordination of numerous regulatory actin-binding proteins (ABPs), many of which bind with affinities that depend on the nucleotide state of actin filament. Cofilin, one of three ABPs that precisely choreograph actin assembly and organization into comet tails that drive motility in vitro, binds and stochastically severs aged ADP actin filament segments of de novo growing actin filaments. Deficiencies in methodologies to track in real time the nucleotide state of actin filaments, as well as cofilin severing, limit the molecular understanding of coupling between actin filament chemical and mechanical states and severing. We engineered a fluorescently labeled cofilin that retains actin filament binding and severing activities. Because cofilin binding depends strongly on the actin-bound nucleotide, direct visualization of fluorescent cofilin binding serves as a marker of the actin filament nucleotide state during assembly. Bound cofilin allosterically accelerates P(i) release from unoccupied filament subunits, which shortens the filament ATP/ADP-P(i) cap length by nearly an order of magnitude. Real-time visualization of filament severing indicates that fragmentation scales with and occurs preferentially at boundaries between bare and cofilin-decorated filament segments, thereby controlling the overall filament length, depending on cofilin binding density.  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