首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
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.
We have used electron paramagnetic probes attached to the ribose of ATP (SL-ATP) to monitor conformational changes in the nucleotide pocket of myosin. Spectra for analogs bound to myosin in the absence of actin showed a high degree of immobilization, indicating a closed nucleotide pocket. In the Actin.Myosin.SL-AMPPNP, Actin.Myosin.SL-ADP.BeF(3), and Actin.Myosin.SL-ADP.AlF(4) complexes, which mimic weakly binding states near the beginning of the power stroke, the nucleotide pocket remained closed. The spectra of the strongly bound Actin.Myosin.SL-ADP complex consisted of two components, one similar to the closed pocket and one with increased probe mobility, indicating a more open pocket, The temperature dependence of the spectra showed that the two conformations of the nucleotide pocket were in equilibrium, with the open conformation more favorable at higher temperatures. These results, which show that opening of the pocket occurs only in the strongly bound states, appear reasonable, as this would tend to keep ADP bound until the end of the power stroke. This conclusion also suggests that force is initially generated by a myosin with a closed nucleotide pocket.  相似文献   

3.
4.
The complexes of pig muscle 3-phosphoglycerate kinase with the substrate MgATP and with the nonsubstrate Mg(2+)-free ATP have been characterized by binding, kinetic, and crystallographic studies. Comparative experiments with ADP and MgADP have also been carried out. In contrast to the less specific and largely ionic binding of Mg(2+)-free ATP and ADP, specific occupation of the adenosine binding pocket by MgATP and MgADP has been revealed by displacement experiments with adenosine and anions, as well as supported by isothermal calorimetric titrations. The Mg(2+)-free nucleotides similarly stabilize the overall protein structure and restrict the conformational flexibility around the reactive thiol groups of helix 13, as observed by differential scanning microcalorimetry and thiol reactivity studies, respectively. The metal complexes, however, behave differently. MgADP, but not MgATP, further increases the conformational stability with respect to its Mg(2+)-free form, which indicates their different modes of binding to the enzyme. Crystal structures of the binary complexes of the enzyme with MgATP and with ATP (2.1 and 1.9 A resolution, respectively) have shown that the orientation and interaction of phosphates of MgATP largely differ not only from those of ATP but also from the previously determined ones of either MgADP [Davies, G. J., Gamblin, S. J., Littlechild, J. A., Dauter, Z., Wilson, K. S., and Watson, H. C. (1994) Acta Crystallogr. D50, 202-209] or the metal complexes of AMP-PNP [May, A., Vas, M., Harlos, K., and Blake, C. C. F. (1996) Proteins 24, 292-303; Auerbach, G., Huber, R., Grattinger, M., Zaiss, K., Schurig, H., Jaenicke, R., and Jacob, U. (1997) Structure 5, 1475-1483] and are more similar to the interactions formed with MgAMP-PCP [Kovári, Z., Flachner, B., Náray-Szabó, G., and Vas, M. (2002) Biochemistry 41, 8796-8806]. Mg(2+) is liganded to both beta- and gamma-phosphates of ATP, while beta-phosphate is linked to the conserved Asp218, i.e., to the N-terminus of helix 8, through a water molecule; the known interactions of either MgADP or the metal complexes of AMP-PNP with the N-terminus of helix 13 and with Asn336 of beta-strand J are absent in the case of MgATP. Fluctuation of MgATP phosphates between two alternative sites has been proposed to facilitate the correct positioning of the mobile side chain of Lys215, and the catalytically competent active site is thereby completed.  相似文献   

5.
The fluorescent probe, 9-anthroylnitrile (ANN), can selectively attach to Ser-180 at the ATP-binding site of subfragment 1 (S1) of skeletal muscle myosin [J. Biol. Chem. 278 (2003) 31891]. We have found that MgATP, MgATPgammaS, MgADP.AlF(4) or MgPP(i), but not MgADP, inhibit the incorporation of ANN into S1. The inhibitory effect of the nucleotide gamma-phosphate group (or its analog) on the modification of S1 with ANN can be explained by the contribution of Ser-180 to the binding of the nucleotide gamma-phosphate at the active site of S1. We have also observed that the incorporation of ANN into S1.MgADP complex is inhibited by actin. These experimental data strongly support the existence of nucleotide-promoted conformational changes revealed by crystal structures of S1 complexes with various nucleotide analogs. They also convincingly show an effect of actin on the environment of Ser-180 at the nucleotide binding site of S1.  相似文献   

6.
The human hypertrophic cardiomyopathy mutation R453C results in one of the more severe forms of the myopathy. Arg-453 is found in a conserved surface loop of the upper 50-kDa domain of the myosin motor domain and lies between the nucleotide binding pocket and the actin binding site. It connects to the cardiomyopathy loop via a long α-helix, helix O, and to Switch-2 via the fifth strand of the central β-sheet. The mutation is, therefore, in a position to perturb a wide range of myosin molecular activities. We report here the first detailed biochemical kinetic analysis of the motor domain of the human β-cardiac myosin carrying the R453C mutation. A recent report of the same mutation (Sommese, R. F., Sung, J., Nag, S., Sutton, S., Deacon, J. C., Choe, E., Leinwand, L. A., Ruppel, K., and Spudich, J. A. (2013) Proc. Natl. Acad. Sci. U.S.A. 110, 12607–12612) found reduced ATPase and in vitro motility but increased force production using an optical trap. Surprisingly, our results show that the mutation alters few biochemical kinetic parameters significantly. The exceptions are the rate constants for ATP binding to the motor domain (reduced by 35%) and the ATP hydrolysis step/recovery stroke (slowed 3-fold), which could be the rate-limiting step for the ATPase cycle. Effects of the mutation on the recovery stroke are consistent with a perturbation of Switch-2 closure, which is required for the recovery stroke and the subsequent ATP hydrolysis.  相似文献   

7.
K Ajtai  L Pótó  T P Burghardt 《Biochemistry》1990,29(33):7733-7741
The nitroxide spin label (iodoacetamido)proxyl (IPSL) was specifically and rigidly attached to sulfhydryl 1 (SH1) on myosin subfragment 1 (S1). The specificity of this label for SH1 was demonstrated by using a technique where the spin label is localized on the electrophoresis-isolated proteolytic fragments of myosin using electron paramagnetic resonance (EPR). Studies of the rigidity of the probe on SH1 indicate that the IPSL is immobilized on the surface of S1 in the presence and absence of the nucleotides MgADP or MgATP. The EPR spectrum of muscle fibers decorated with IPSL-S1 shows that the IPSL-S1 rotates from its orientation in rigor upon binding MgADP. The angular displacement due to nucleotide binding is larger than that detected with the (maleimido)tempo spin label [Ajtai, K., French, A. R., & Burghardt, T. P. (1989) Biophys. J. 56, 535-541], demonstrating that the IPSL is oriented on the myosin cross-bridge in a manner that is favorable for detecting cross-bridge rotation during the rigor to MgADP state transition.  相似文献   

8.
Hiratsuka T 《Biochemistry》2006,45(4):1234-1241
The fluorescent probe 3-[4-(3-phenyl-2-pyrazolin-1-yl)benzene-1-sulfonyl amido]phenylboronic acid (PPBA) acts as a fluorescent inhibitor for the ATPases of skeletal [Hiratsuka (1994) J. Biol. Chem. 269, 27251-27257] and Dictyostelium discoideum [Bobkov et al. (1997) J. Muscle Res. Cell Motil. 18, 563-571] myosins. The former paper suggested that, upon addition of excess nucleotides to the binary complex of subfragment-1 from skeletal myosin (S1) with PPBA, a stable ternary complex of S1 with PPBA and nucleotide is formed. Useful fluorescence properties of PPBA enable us to distinguish the conformation of the myosin ATPase at the ATP state from that at the ADP state. In the present paper, to determine the PPBA-binding site in the complexes, enzymatic and fluorescence properties of the S1.PPBA.nucleotide complexes were investigated. Upon formation of the ternary complex with ATP, a new peak appeared at 398 nm in the PPBA fluorescence spectrum. Experiments using model compounds of aromatic amino acid suggested that this fluorescence peak at 398 nm is originated from PPBA interacting with Phe residue(s). Taking into account differences in fluorescence spectra between complexes of S1 and those of subfragment-1 from D. discoideum myosin (S1dC), in the ternary complex of S1 formed with ATP, PPBA was suggested to interact with Phe residue(s) that is absent in S1dC. Docking simulation of PPBA on the S1.nucleotide complex revealed that Phe472 interacts with PPBA. Binding sites of PPBA and blebbistatin, an inhibitor showing high affinity and selectivity toward myosin II [Kovács et al. (2004) J. Biol. Chem. 279, 35557-35563], seem to overlap at least partly.  相似文献   

9.
Chen X  Grammer J  Lawson JD  Cooke R  Pate E  Yount RG 《Biochemistry》2002,41(8):2609-2620
The photoaffinity spin-labeled non-nucleoside ATP analogue, 2-(4-azido-2-nitrophenyl)amino-2,2-(1-oxyl-2,2,6,6-tetramethyl-4-piperidylidene)di(oxymethylene)ethyl triphosphate (SSL-NANTP), has been shown to be a substrate for skeletal mysoin subfragment 1 (S1) that can be photoincorporated at the active site of S1 [Chen, X., et al. (2000) Bioconjugate Chem. 11, 725-733]. Electron paramagnetic resonance spectroscopy shows that the probe undergoes restricted motion with respect to the protein. The parent compound, NANTP (2-[(4-azido-2-nitrophenyl)amino]ethyl triphosphate), is specifically photoincorporated at Trp-130 on the amino-terminal 23 kDa tryptic fragment in rabbit skeletal myosin. Surprisingly, amino acid sequence analysis shows that SSL-NANTP is photoincorporated on the carboxy-terminal 20 kDa tryptic fragment at Lys-681 on the side opposite Trp-130 in the nucleotide pocket. This is the first direct evidence showing that this residue in the 20 kDa tryptic fragment is close enough to the active site to be photolabeled by trapped ATP analogues. After actin treatment in the presence of MgATP, SSL-NANDP-labeled myosin S1 had normal ATPase activity, indicating that photolabeling did not significantly alter the enzymatic properties of S1. Photoincorporated SSL-NANDP was bound inside the nucleotide site of S1, with an effective concentration of 20 mM as judged by the concentration of MgADP needed to displace it. Molecular dynamics simulations suggest that the ability of NANTP and SSL-NANTP to photolabel different sites results from different orientations of the phenyl ring in the active site. For SSL-NANTP, the p-azido group on the phenyl ring points toward Lys-681. For NANTP, it points in the opposite direction toward Trp-130.  相似文献   

10.
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   

11.
We have previously shown that inhibition of the ATPase activity of skeletal muscle myosin subfragment 1 (S1) by caldesmon is correlated with the inhibition of S1 binding in the presence of ATP or pyrophosphate (Chalovich, J., Cornelius, P., and Benson, C. (1987) J. Biol Chem. 262, 5711-5716). In contrast, Lash et al. (Lash, J., Sellers, J., and Hathaway, D. (1986) J. Biol. Chem. 261, 16155-16160) have shown that the inhibition of ATPase activity of smooth muscle heavy meromyosin (HMM) by caldesmon is correlated with an increase in the binding of HMM to actin in the presence of ATP. We now show, in agreement, that caldesmon does increase the binding of smooth muscle HMM to actin-tropomyosin while decreasing the ATPase activity. The effect of caldesmon on the binding of smooth HMM is reversed by Ca2+-calmodulin. Caldesmon strengthens the binding of smooth S1.ATP and skeletal HMM.ATP to actin-tropomyosin but to a lesser extent than smooth HMM.ATP. Furthermore, this increase in binding of smooth S1.ATP and skeletal HMM.ATP does not parallel the inhibition of ATPase activity. In contrast, in the absence of ATP, all smooth and skeletal myosin subfragments compete with caldesmon for binding to actin. Thus, the effect that caldesmon has on the binding of myosin subfragments to actin-tropomyosin depends on the source of myosin, the type of subfragment, and the nucleotide present. The inhibition of actin-activated ATP hydrolysis by caldesmon, however, is not greatly different for different smooth and skeletal myosin subfragments. Evidence is presented that caldesmon inhibits actin-activated ATP hydrolysis by attenuating the productive interaction between myosin and actin that normally accelerates ATP hydrolysis. The increased binding seen by some myosin subfragments, in the presence of ATP, may be due to binding of these subfragments to a nonproductive site on actin-caldesmon. The subfragments which show an increase in binding in the presence of ATP and caldesmon appear to bind directly to caldesmon as demonstrated by affinity chromatography.  相似文献   

12.
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   

13.
M M Werber  Y M Peyser  A Muhlrad 《Biochemistry》1992,31(31):7190-7197
Beryllium and aluminum fluorides are good phosphate analogues. These compounds, like orthovanadate, form stable complexes with myosin subfragment 1 (S1) in the presence of MgADP. The formation of the stable S1-nucleotide complexes is characterized by the loss of ATPase activity. For the complete loss of ATPase activity there was necessary a higher concentration of aluminum than of beryllium or vanadate. In the presence of MgATP the onset of the inhibition is delayed, which indicates that stable complexes cannot form when a specific site is occupied by the gamma-phosphate of ATP or by P(i) derived from the gamma-phosphate. The half-lives of the S1-MgADP-(BeF3-), S1-MgADP-(AlF4-), and S1-MgADP-Vi complexes at 0 degrees C are 7, 2, and 4 days, respectively. In the presence of actin the rate of decomposition of all of the complexes is significantly enhanced; however, the order of decomposition is reversed, the fastest rate being observed with beryllium and the slowest with aluminum. The formation of the S1-MgADP-(BeF3-) and S1-MgADP-(AlF4-) complexes is accompanied by an increase in tryptophan fluorescence similar to that observed upon addition of MgATP to S1. The fluorescence increase develops rather slowly, by suggesting that the rate-limiting step in the formation of the stable complex is an isomerization. The rate of the fluorescence change accompanying the formation of the Be complex is faster than that for the Al complex. Addition of vanadate to S1 causes a static quenching of the tryptophan fluorescence.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Combined mutation of "catalytic carboxylates" in both nucleotide binding domains (NBDs) of P-glycoprotein generates a conformation capable of tight binding of 8-azido-ADP (Sauna, Z. E., Müller, M., Peng, X. H., and Ambudkar, S. V. (2002) Biochemistry 41, 13989-14000). Here we characterized this conformation using pure mouse MDR3 P-glycoprotein and natural MgATP and MgADP. Mutants E552A/E1197A, E552Q/E1197Q, E552D/E1197D, and E552K/E1197K had low but real ATPase activity in the order Ala > Gln > Asp > Lys, emphasizing the requirement for Glu stereochemistry. Mutant E552A/E1197A bound MgATP and MgADP (1 mol/mol) with K(d) 9.2 and 92 microm, showed strong temperature sensitivity of MgATP binding and equal dissociation rates for MgATP and MgADP. With MgATP as the added ligand, 80% of bound nucleotide was in the form of ATP. None of these parameters was vanadate-sensitive. The other mutants showed lower stoichiometry of MgATP and MgADP binding, in the order Ala > Gln > Asp > Lys. We conclude that the E552A/E1197A mutation arrests the enzyme in a conformation, likely a stabilized NBD dimer, which occludes nucleotide, shows preferential binding of ATP, does not progress to a normal vanadate-sensitive transition state, but hydrolyzes ATP and releases ADP slowly. Impairment of turnover is primarily due to inability to form the normal transition state rather than to slow ADP release. The Gln, Asp, and Lys mutants are less effective at stabilizing the occluded nucleotide, putative dimeric NBD, conformation. We envisage that in wild-type the occluded nucleotide conformation occurs transiently after MgATP binds to both NBDs with associated dimerization, and before progression to the transition state.  相似文献   

15.
Polarized fluorimetry technique and ghost muscle fibers containing tropomyosin were used to study effects of caldesmon (CaD) and recombinant peptides CaDH1 (residues 506-793), CaDH2 (residues 683-767), CaDH12 (residues 506-708) and 658C (residues 658-793) on the orientation and mobility of fluorescent label 1.5-IAEDANS specifically bound to Cys-707 of myosin subfragment-1 (S1) in the absence of nucleotide, and in the presence of MgADP, MgAMP-PNP, MgATPgammaS or MgATP. It was shown that at modelling different intermediates of actomyosin ATPase, the orientation and mobility of dye dipoles changed discretely, suggesting a multi-step changing of the myosin head structural state in ATP hydrolysis cycle. The maximum difference in orientation and mobility of the oscillator (4 degrees and 30%, respectively) was observed between actomyosin in the presence of MgATP, and actomyosin in the presence of MgADP. Caldesmon actin-binding sites C and B' inhibit formation of actomyosin strong binding states, while site B activates it. It is suggested that actin-myosin interaction in ATP hydrolysis cycle initiates nucleotide-dependent rotation of myosin motor domain, or that of its site for dye binding as well as the change in myosin head mobility. Caldesmon drives ATP hydrolysis cycle by shifting the equilibrium between strong and weak forms of actin-myosin binding.  相似文献   

16.
Atomic structures of scallop myosin subfragment 1(S1) with the bound MgADP, MgAMPPNP, and MgADP.BeF(x) provide crystallographic evidence for a destabilization of the helix containing reactive thiols SH1 (Cys703) and SH2 (Cys693). A destabilization of this helix was not observed in previous structures of S1 (from chicken skeletal, Dictyostelium discoideum, and smooth muscle myosins), including complexes for which solution experiments indicated such a destabilization. In this study, the factors that influence the SH1-SH2 helix in scallop S1 were examined using monofunctional and bifunctional thiol reagents. The rate of monofunctional labeling of scallop S1 was increased in the presence of MgADP and MgATPgammaS but was inhibited by MgADP.V(i) and actin. The resulting changes in ATPase activities of S1 were symptomatic of SH2 and not SH1 modification, which was confirmed by mass spectrometry analysis. With bifunctional reagents of various lengths, cross-linking did not occur on a short time scale in the absence of nucleotides. In the presence of MgADP, cross-linking was greatly enhanced for all of the reagents. These reactions, as well as the formation of a disulfide bond between SH1 and SH2, were much faster in scallop S1.ADP than in rabbit skeletal S1.ADP and were rate-limited by the initial attachment of the reagent to scallop S1. The cross-linking sites were mapped by mass spectrometry to SH1 and SH2. These results reveal isoform-specific differences in the conformation and dynamics of the SH1-SH2 helix, providing a possible explanation for destabilization of this helix in some scallop S1 but not in other S1 isoform structures.  相似文献   

17.
In our previous study [Chalovich, J. M., Greene, L. E., & Eisenberg, E. (1983) Proc. Natl. Acad. Sci. U.S.A. 80, 4909-4913], myosin subfragment 1 that was modified by having its two reactive thiol groups cross-linked by N,N'-p-phenylenedimaleimide (pPDM) was found to resemble the myosin subfragment 1-adenosine 5'-triphosphate (S-1.ATP) complex in its interaction with actin. In the present study, we examined the effect of actin on adenosine 5'-diphosphate (ADP) trapped at the active site of pPDM.S-1. Our results indicate first that, in the presence of actin, ADP is no longer trapped at the active site but exchanges rapidly with free nucleotide. Different pPDM.S-1.nucleotide complexes were then formed by exchanging nucleotide into the active site of pPDM.S-1 in the presence of actin. The binding of pPDM.S-1.ATP or pPDM.S-1.PPi to actin is virtually identical with that of unmodified S-1 in the presence of ATP. Specifically, at mu = 18 mM, 25 degrees C, pPDM.S-1.ATP or pPDM.S-1.PPi binds to unregulated actin with the same affinity as does S-1.ATP, and this binding does not appear to be affected by troponin-tropomyosin. On the other hand, pPDM.S-1.ADP and pPDM.S-1 with no bound nucleotide both show a small, but significant, difference between their binding to actin and the binding of S-1.ATP; pPDM.S-1 and pPDM.S-1.ADP both bind about 2- to 3-fold more strongly to unregulated actin than does S-1.ATP.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
3'-O-(4-Benzoyl)benzoyl-ATP (Bz2ATP), an analog of ATP containing a photoreactive benzophenone moiety, was used as a probe of the ATP binding site of myosin subfragment 1 (SF1). The inactivation of SF1 NH+4-EDTA ATPase by the bifunctional thiol crosslinking system cobalt(II)/cobalt(III) phenanthroline complexes was enhanced by Bz2ATP to the same degree as by ATP. This treatment resulted in the stable trapping of Bz2ATP at the active site in nearly stoichiometric amounts in a manner exactly analogous to ATP (Wells, J.A., and Yount, R.G. (1979) Proc. Natl. Acad. Sci. U.S.A. 76, 4966-4970). Irradiation of SF1 containing trapped [3H]Bz2ATP gave approximately 50% covalent incorporation of the trapped nucleotide into the enzyme. Analysis of photolabeled SF1 by gel electrophoresis showed that all of the [3H]Bz2ATP was attached to the 95-kDa heavy chain fragment. No label was found in the light chains. Similar analysis of the same protein after limited trypsin treatment demonstrated that approximately 75% of the [3H]Bz2ATP was bound to the central 50-kDa peptide and its 75-kDa precursor from the heavy chain. The N-terminal 25-kDa tryptic peptide, shown to be photolabeled by other ATP analogs (Szilagyi, L., Balint, M., Sreter, F.A., and Gergely, J. (1979) Biochem. Biophys. Res. Commun. 87, 936-945; Okamoto, Y., and Yount, R.G. (1983) Biophys. J. 41, 298a), was not labeled (less than 1%) by Bz2ATP. These results demonstrate that portions of the 50 kDa-peptide of the heavy chain are within 6-7 A of the ATP binding site on SF1 and possibly contribute to nucleotide binding.  相似文献   

19.
Stability constants for the Mg2+ and Cd2+ complexes of ATP, ADP, ATP alpha S, ATP beta S, and ADP alpha S have been determined at 30 degrees C and mu = 0.1 M by 31P NMR. Besides being of the utmost importance for determining species distributions for enzymatic studies, these constants allow an estimation of the preference of Cd2+ for sulfur vs. oxygen coordination in phosphorothioate complexes. Stability constants for Mg2+ complexes decreases when sulfur replaces oxygen (log K: ADP, 4.11; ADP alpha S, 3.66; ATP, 4.70; ATP alpha S, 4.47; ATP beta S, 4.04) because of (a) a statistical factor resulting from the loss of one potential phosphate oxygen ligand and (b) either an alteration in the charge distribution between oxygen and sulfur or destabilization of the chelate ring structure by loss of an internal hydrogen bond between an oxygen of coordinated phosphate and metal-bound water. Cd2+ complexes with sulfur-substituted nucleotides are more stable than those without sulfur (log K: ADP, 3.58; ADP alpha S, 4.95; ATP, 4.36; ATP alpha S, 4.42; ATP beta S, 5.44) because of the preferential binding of Cd2+ to sulfur rather than oxygen, which we estimate to be approximately 60 in CdADP alpha S and CdATP beta S. The proportion of tridentate coordination is estimated to be 50-60% in MgATP and MgATP beta S, approximately 27% in MgATP alpha S, approximately 16% in CdATP or CdATP beta S, but approximately 75% in CdATP alpha S. By analysis of the data of Jaffe and Cohn [Jaffe, E. K., & Cohn, M. (1979) J. Biol. Chem. 254, 10839], we conclude that the preference for oxygen over sulfur coordination to ATP beta S is 31 000 for Mg2+, 3100-3900 for Ca2+, and 158-193 for Mn2+. Proton NMR demonstrates that bidentate Cd2+ complexes form intramolecular chelates with the N-7 of adenine while Mg2+ nucleotides and the tridenate CdATP alpha S do not. An analysis of the 31P NMR line widths shows that the rate constants for dissociation of MgADP and MgATP are both 7000 s-1 while the association rate constants are 7 X 10(7) and 4 X 10(8) M-1 s-1, respectively. The observed dependence of the line width on nucleotide concentration is best explained by a base-stacking model at nucleotide concentrations above 5 mM.  相似文献   

20.
KATP channels were reconstituted in COSm6 cells by coexpression of the sulfonylurea receptor SUR1 and the inward rectifier potassium channel Kir6.2. The role of the two nucleotide binding folds of SUR1 in regulation of KATP channel activity by nucleotides and diazoxide was investigated. Mutations in the linker region and the Walker B motif (Walker, J.E., M.J. Saraste, M.J. Runswick, and N.J. Gay. 1982. EMBO [Eur. Mol. Biol. Organ.] J. 1:945–951) of the second nucleotide binding fold, including G1479D, G1479R, G1485D, G1485R, Q1486H, and D1506A, all abolished stimulation by MgADP and diazoxide, with the exception of G1479R, which showed a small stimulatory response to diazoxide. Analogous mutations in the first nucleotide binding fold, including G827D, G827R, and Q834H, were still stimulated by diazoxide and MgADP, but with altered kinetics compared with the wild-type channel. None of the mutations altered the sensitivity of the channel to inhibition by ATP4−. We propose a model in which SUR1 sensitizes the KATP channel to ATP inhibition, and nucleotide hydrolysis at the nucleotide binding folds blocks this effect. MgADP and diazoxide are proposed to stabilize this desensitized state of the channel, and mutations at the nucleotide binding folds alter the response of channels to MgADP and diazoxide by altering nucleotide hydrolysis rates or the coupling of hydrolysis to channel activation.  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