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.
To understand the molecular mechanism underlying phosphoryl transfer of cAMP-dependent protein kinase, the structure of the catalytic subunit in complex with ADP, aluminum fluoride, Mg2+ ions and a substrate peptide was determined at 2.0 A resolution. Aluminum fluoride was modeled as AlF3 in a planar geometry; it is positioned 2.3 A from both the donor oxygen of ADP and the hydroxyl group of the recipient Ser residue. In this configuration, the aluminum atom forms a trigonal bipyramidal coordination with the oxygen atoms of the donor and recipient groups at the apical positions. This arrangement suggests that aluminum fluoride mimics the transition state and provides the first direct structural evidence for the in-line mechanism of phosphoryl transfer in a protein kinase.  相似文献   

2.
Phosphoenolpyruvate carboxykinase (PCK) catalyzes the conversion of oxaloacetate (OAA) to PEP and carbon dioxide with the subsequent conversion of nucleoside triphosphate to nucleoside diphosphate (NDP). The 1.9 A resolution structure of Escherichia coli PCK consisted of a 275-residue N-terminal domain and a 265-residue C-terminal domain with the active site located in a cleft between these domains. Each domain has an alpha/beta topology and the overall structure represents a new protein fold. Furthermore, PCK has a unique mononucleotide-binding fold. The 1.8 A resolution structure of the complex of ATP/Mg(2+)/oxalate with PCK revealed a 20 degrees hinge-like rotation of the N- and C-terminal domains, which closed the active site cleft. The ATP was found in the unusual syn conformation as a result of binding to the enzyme. Along with the side chain of Lys254, Mg(2+) neutralizes charges on the P beta and P gamma oxygen atoms of ATP and stabilizes an extended, eclipsed conformation of the P beta and P gamma phosphoryl groups. The sterically strained high-energy conformation likely lowers the free energy of activation for phosphoryl transfer. Additionally, the gamma-phosphoryl group becomes oriented in-line with the appropriate enolate oxygen atom, which strongly supports a direct S(N)2-type displacement of this gamma-phosphoryl group by the enolate anion. In the 2.0 A resolution structure of the complex of PCK/ADP/Mg(2+)/AlF(3), the AlF(3) moiety represents the phosphoryl group being transferred during catalysis. There are three positively charged groups that interact with the fluorine atoms, which are complementary to the three negative charges that would occur for an associative transition state.  相似文献   

3.
Phosphoribosylaminoimidazole-succinocarboxamide synthetase (SAICAR synthetase) converts 4-carboxy-5-aminoimidazole ribonucleotide (CAIR) to 4-(N-succinylcarboxamide)-5-aminoimidazole ribonucleotide (SAICAR). The enzyme is a target of natural products that impair cell growth. Reported here are the crystal structures of the ADP and the ADP.CAIR complexes of SAICAR synthetase from Escherichia coli, the latter being the first instance of a CAIR-ligated SAICAR synthetase. ADP and CAIR bind to the active site in association with three Mg(2+), two of which coordinate the same oxygen atom of the 4-carboxyl group of CAIR; whereas, the third coordinates the alpha- and beta-phosphoryl groups of ADP. The ADP.CAIR complex is the basis for a transition state model of a phosphoryl transfer reaction involving CAIR and ATP, but also supports an alternative chemical pathway in which the nucleophilic attack of l-aspartate precedes the phosphoryl transfer reaction. The polypeptide fold for residues 204-221 of the E. coli structure differs significantly from those of the ligand-free SAICAR synthetase from Thermatoga maritima and the adenine nucleotide complexes of the synthetase from Saccharomyces cerevisiae. Conformational differences between the E. coli, T. maritima, and yeast synthetases suggest the possibility of selective inhibition of de novo purine nucleotide biosynthesis in microbial organisms.  相似文献   

4.
The crystal structures of two complexes of dethiobiotin synthetase, enzyme-diaminopelargonic acid-MgADP-AlF3 and enzyme-dethiobiotin-MgADP-Pi, respectively, have been determined to 1.8 A resolution. In dethiobiotin synthetase, AlF3 together with carbamylated diaminopelargonic acid mimics the phosphorylated reaction intermediate rather than the transition state complex for phosphoryl transfer. Observed differences in the binding of substrate, diaminopelargonic acid, and the product, dethiobiotin, suggest considerable displacements of substrate atoms during the ring closure step of the catalytic reaction. In both complexes, two metal ions are observed at the active site, providing evidence for a two-metal mechanism for this enzyme.  相似文献   

5.
Eukaryotic protein kinases catalyze the phosphoryl transfer of the gamma-phosphate of ATP to the serine, threonine, or tyrosine residue of protein substrates. The catalytic mechanism of phospho-CDK2/cyclin A (pCDK2/cyclin A) has been probed with structural and kinetic studies using the trigonal NO(3)(-) ion, which can be viewed as a mimic of the metaphosphate transition state. The crystal structure of pCDK2/cyclin A in complex with Mg(2+)ADP, nitrate, and a heptapeptide substrate has been determined at 2.7 A. The nitrate ion is located between the beta-phosphate of ADP and the hydroxyl group of the serine residue of the substrate. In one molecule of the asymmetric unit, the nitrate is close to the beta-phosphate of ADP (distance from the nitrate nitrogen to the nearest beta-phosphate oxygen of 2.5 A), while in the other subunit, the nitrate is closer to the substrate serine (distance of 2.1 A). Kinetic studies demonstrate that nitrate is not an effective inhibitor of protein kinases, consistent with the structural results that show the nitrate ion makes few stabilizing interactions with CDK2 at the catalytic site. The binding of orthovanadate was also investigated as a mimic of a pentavalent phosphorane intermediate of an associative mechanism for phosphoryl transfer. No vanadate was observed bound in a 3.4 A resolution structure of pCDK2/cyclin A in the presence of Mg(2+)ADP, and vanadate did not inhibit the kinase reaction. The results support the notion that the protein kinase reaction proceeds through a mostly dissociative mechanism with a trigonal planar metaphosphate intermediate rather than an associative mechanism that involves a pentavalent phosphorane intermediate.  相似文献   

6.
To study the catalytic mechanism of phosphorylation catalyzed by cAMP-dependent protein kinase (PKA) a structure of the enzyme-substrate complex representing the Michaelis complex is of specific interest as it can shed light on the structure of the transition state. However, all previous crystal structures of the Michaelis complex mimics of the PKA catalytic subunit (PKAc) were obtained with either peptide inhibitors or ATP analogs. Here we utilized Ca2+ ions and sulfur in place of the nucleophilic oxygen in a 20-residue pseudo-substrate peptide (CP20) and ATP to produce a close mimic of the Michaelis complex. In the ternary reactant complex, the thiol group of Cys-21 of the peptide is facing Asp-166 and the sulfur atom is positioned for an in-line phosphoryl transfer. Replacement of Ca2+ cations with Mg2+ ions resulted in a complex with trapped products of ATP hydrolysis: phosphate ion and ADP. The present structural results in combination with the previously reported structures of the transition state mimic and phosphorylated product complexes complete the snapshots of the phosphoryl transfer reaction by PKAc, providing us with the most thorough picture of the catalytic mechanism to date.  相似文献   

7.
N-Acetyl-L-glutamate kinase (NAGK), the structural paradigm of the enzymes of the amino acid kinase family, catalyzes the phosphorylation of the gamma-COO(-) group of N-acetyl-L-glutamate (NAG) by ATP. We determine here the crystal structures of NAGK complexes with MgADP, NAG and the transition-state analog AlF(4)(-); with MgADP and NAG; and with ADP and SO(4)(2-). Comparison of these structures with that of the MgAMPPNP-NAG complex allows to delineate three successive steps during phosphoryl transfer: at the beginning, when the attacking and leaving O atoms and the P atom are imperfectly aligned and the distance between the attacking O atom and the P atom is 2.8A; midway, at the bipyramidal intermediate, with nearly perfect alignment and a distance of 2.3A; and, when the transfer is completed. The transfer occurs in line and is strongly associative, with Lys8 and Lys217 stabilizing the transition state and the leaving group, respectively, and with Lys61, in contrast with an earlier proposal, not being involved. Three water molecules found in all the complexes play, together with Asp162 and the Mg, crucial structural roles. Two glycine-rich loops (beta1-alphaA and beta2-alphaB) are also very important, moving in the different complexes in concert with the ligands, to which they are hydrogen-bonded, either locking them in place for reaction or stabilizing the transition state. The active site is too narrow to accommodate the substrates without compressing the reacting groups, and this compressive strain appears a crucial component of the catalytic mechanism of NAGK, and possibly of other enzymes of the amino acid kinase family such as carbamate kinase. Initial binding of the two substrates would require a different enzyme conformation with a wider active site, and the energy of substrate binding would be used to change the conformation of the active center, causing substrate strain towards the transition state.  相似文献   

8.
Davies DR  Hol WG 《FEBS letters》2004,577(3):315-321
The formation of transition state mimics of phosphoryl transfer reactions with the metal oxoanion vanadate is a powerful technique in macromolecular crystallography. The tendency of vanadate to form pentacovalent complexes exhibiting trigonal bipyramidal geometry makes this compound a close approximation of the transition state for such reactions. In many cases, vanadate complexes provide the most accurate visualization of the transition state that can be reasonably achieved. A survey of the Protein Data Bank reveals that a relatively small number of structures (39, representing 23 unique proteins) include vanadate, yet these structures represent four of the six E.C. categories of enzymes, and were obtained in crystals with pH values ranging from 5.0 to 7.8. Vanadate has additional advantages over other compounds such as aluminum fluoride, beryllium fluoride and nitrate used for visualization of transition state mimics in that vanadate readily forms covalent bonds with a variety of ligands and has produced a wider variety of transition state mimics. Given the hundreds of crystal structures that have been solved for phosphoryl transfer enzymes, it is surprising that vanadate has not been used more frequently for visualization of transition state analogs. We propose that an opportunity exists for vanadate to become a more commonly utilized component of the macromolecular crystallographer's toolbox.  相似文献   

9.
Aluminum fluoride has become an important tool for investigating the mechanism of phosphoryl transfer, an essential reaction that controls a host of vital cell functions. Planar AlF(3) or AlF(4)(-) molecules are proposed to mimic the phosphoryl group in the catalytic transition state. Acetate kinase catalyzes phosphoryl transfer of the ATP gamma-phosphate to acetate. Here we describe the inhibition of acetate kinase from Methanosarcina thermophila by preincubation with MgCl(2), ADP, AlCl(3), NaF, and acetate. Preincubation with butyrate in place of acetate did not significantly inhibit the enzyme. Several NTPs can substitute for ATP in the reaction, and the corresponding NDPs, in conjunction with MgCl(2), AlCl(3), NaF, and acetate, inhibit acetate kinase activity. Fluorescence quenching experiments indicated an increase in binding affinity of acetate kinase for MgADP in the presence of AlCl(3), NaF, and acetate. These and other characteristics of the inhibition indicate that the transition state analog, MgADP-aluminum fluoride-acetate, forms an abortive complex in the active site. The protection from inhibition by a non-hydrolyzable ATP analog or acetylphosphate, in conjunction with the strict dependence of inhibition on the presence of both ADP and acetate, supports a direct in-line mechanism for acetate kinase.  相似文献   

10.
D Herschlag  W P Jencks 《Biochemistry》1990,29(21):5172-5179
The second-order rate constants for reaction of the Mg2+ complexes of phosphorylated pyridine monoanions with Mg(OH)+ are 10(4)-10(6)-fold larger than the second-order rate constants for their reaction with water (25 degrees C, ionic strength 1.5). Of the 10(6)-fold rate enhancement with the phosphorylated 4-morpholinopyridine/Mg2 complex, approximately 10(4)-fold is attributed to the greater nucleophilicity of Mg(OH)+ compared with water. The remaining catalysis of approximately 10(2)-fold is attributed to induced intramolecularity from positioning of the hydroxide ion and phosphoryl group by the Mg2+ ions. This reaction may provide a model for the role of a metal ion in increasing the concentration of the anions of enolpyruvate and serine and holding the nucleophile in the correct position for phosphoryl transfer in the reactions catalyzed by pyruvate kinase and alkaline phosphatase, for example. Some mechanisms that can provide catalysis of phosphoryl transfer through a metaphosphate-like transition state are reviewed briefly.  相似文献   

11.
The paper reports a study of the kinetics of the reaction between phosphoenolpyruvate, ADP and Mg(2+) catalysed by yeast pyruvate kinase when activated by fructose 1,6-diphosphate and K(+). The experimental results indicate that the reaction mechanism is of the Ordered Tri Bi type with the substrates binding in the order phosphoenolpyruvate, ADP and Mg(2+). Direct phosphoryl transfer takes place in the quaternary complex, with pyruvate released before MgATP. A dead-end enzyme-pyruvate complex is also indicated. Values have been determined for the Michaelis, dissociation and inhibition constants of the reaction. Several of the rate constants involved have also been evaluated.  相似文献   

12.
Acetate kinase catalyzes transfer of the gamma-phosphate of ATP to acetate. The only crystal structure reported for acetate kinase is the homodimeric enzyme from Methanosarcina thermophila containing ADP and sulfate in the active site (Buss, K. A., Cooper, D. C., Ingram-Smith, C., Ferry, J. G., Sanders, D. A., and Hasson, M. S. (2001) J. Bacteriol. 193, 680-686). Here we report two new crystal structure of the M. thermophila enzyme in the presence of substrate and transition state analogs. The enzyme co-crystallized with the ATP analog adenosine 5'-[gamma-thio]triphosphate contained AMP adjacent to thiopyrophosphate in the active site cleft of monomer B. The enzyme co-crystallized with ADP, acetate, Al(3+), and F(-) contained a linear array of ADP-AlF(3)-acetate in the active site cleft of monomer B. Together, the structures clarify the substrate binding sites and support a direct in-line transfer mechanism in which AlF(3) mimics the meta-phosphate transition state. Monomers A of both structures contained ADP and sulfate, and the active site clefts were closed less than in monomers B, suggesting that domain movement contributes to catalysis. The finding that His(180) was in close proximity to AlF(3) is consistent with a role for stabilization of the meta-phosphate that is in agreement with a previous report indicating that this residue is essential for catalysis. Residue Arg(241) was also found adjacent to AlF(3), consistent with a role for stabilization of the transition state. Kinetic analyses of Arg(241) and Arg(91) replacement variants indicated that these residues are essential for catalysis and also indicated a role in binding acetate.  相似文献   

13.
14.
BACKGROUND: Thymidylate kinase (TMPK) is a nucleoside monophosphate kinase that catalyzes the reversible phosphoryltransfer between ATP and TMP to yield ADP and TDP. In addition to its vital role in supplying precursors for DNA synthesis, human TMPK has an important medical role participating in the activation of a number of anti-HIV prodrugs. RESULTS: Crystal structures of human TMPK in complex with TMP and ADP, TMP and the ATP analog AppNHp, TMP with ADP and the phosphoryl analog AlF(3), TDP and ADP, and the bisubstrate analog TP(5)A were determined. The conformations of the P-loop, the LID region, and the adenine-binding loop vary according to the nature of the complex. Substitution of ADP by AppNHp results in partial closure of the P-loop and the rotation of the TMP phosphate group to a catalytically unfavorable position, which rotates back in the AlF(3) complex to a position suitable for in-line attack. In the fully closed state observed in the TP(5)A and the TDP-ADP complexes, Asp15 interacts strongly with the 3'-hydroxyl group of TMP. CONCLUSIONS: The observed changes of nucleotide state and conformation and the corresponding protein structural changes are correlated with intermediates occurring along the reaction coordinate and show the sequence of events occurring during phosphate transfer. The low catalytic activity of human TMPK appears to be determined by structural changes required to achieve catalytic competence and it is suggested that a mechanism might exist to accelerate the activity.  相似文献   

15.
The conformation of di- and triphosphate nucleosides in the active site of ATPsynthase (H(+)-ATPase) from thermophilic Bacillus PS3 (TF1) and their interaction with Mg(2+)/Mn(2+) cations have been investigated using EPR, ESEEM, and HYSCORE spectroscopies. For a ternary complex formed by a stoichiometric mixture of TF1, Mn(2+), and ADP, the ESEEM and HYSCORE data reveal a (31)P hyperfine interaction with Mn(2+) (|A((31)P)| approximately 5.20 MHz), significantly larger than that measured for the complex formed by Mn(2+) and ADP in solution (|A((31)P)| approximately 4.50 MHz). The Q-band EPR spectrum of the Mn.TF1.ADP complex indicates that the Mn(2+) binds in a slightly distorted environment with |D| approximately 180 x 10(-4) cm(-1) and |E| approximately 50 x 10(-4) cm(-1). The increased hyperfine coupling with (31)P in the presence of TF1 reflects the specific interaction between the central Mn(2+) and the ADP beta-phosphate, illustrating the role of the enzyme active site in positioning the phosphate chain of the substrate for efficient catalysis. Results with the ternary Mn.TF1.ATP and Mn.TF1.AMP-PNP complexes are interpreted in a similar way with two hyperfine couplings being resolved for each complex (|A((31)P(beta))| approximately 4.60 MHz and |A((31)P(gamma))| approximately 5.90 MHz with ATP, and |A((31)P(beta))| approximately 4.20 MHz and |A((31)P(gamma))| approximately 5.40 MHz with AMP-PNP). In these complexes, the increased hyperfine coupling with (31)P(gamma) compared with (31)P(beta) reflects the smaller Mn.P distance with the gamma-phosphate compared with the beta-phosphate as found in the crystal structure of the analogous enzyme from mitochondria [3.53 vs 3.70 A (Abrahams, J. P., Leslie, A. G. W., Lutter, R., and Walker, J. E. (1994) Nature 370, 621-628)] and the different binding modes of the two phosphate groups. The ESEEM and HYSCORE data of a complex formed with Mn(2+), ATP, and the isolated beta subunit show that the (31)P hyperfine coupling is close to that measured in the absence of the protein, indicating a poorly structured nucleotide site in the isolated beta subunit in the presence of ATP. The inhibition data obtained for TF1 incubated in the presence of Mg(2+), ADP, Al(NO(3))(3), and NaF indicate the formation of the inhibited complex with the transition state analogue namely Mg.TF1.ADP.AlF(x) with the equilibrium dissociation constant K(D) = 350 microM and rate constant k = 0.02 min(-1). The ESEEM and HYSCORE data obtained for an inhibited TF1 sample, Mn.TF1.ADP.AlF(x), confirm the formation of the transition state analogue with distinct spectroscopic footprints that can be assigned to Mn.(19)F and Mn.(27)Al hyperfine interactions. The (31)P(beta) hyperfine coupling that is measured in the inhibited complex with the transition state analogue (|A((31)P(beta))| approximately 5.10 MHz) is intermediate between those measured in the presence of ADP and ATP and suggests an increase in the bond between Mn and the P(beta) from ADP upon formation of the transition state.  相似文献   

16.
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   

17.
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   

18.
Productive cis folding by the chaperonin GroEL is triggered by the binding of ATP but not ADP, along with cochaperonin GroES, to the same ring as non-native polypeptide, ejecting polypeptide into an encapsulated hydrophilic chamber. We examined the specific contribution of the gamma-phosphate of ATP to this activation process using complexes of ADP and aluminium or beryllium fluoride. These ATP analogues supported productive cis folding of the substrate protein, rhodanese, even when added to already-formed, folding-inactive cis ADP ternary complexes, essentially introducing the gamma-phosphate of ATP in an independent step. Aluminium fluoride was observed to stabilize the association of GroES with GroEL, with a substantial release of free energy (-46 kcal/mol). To understand the basis of such activation and stabilization, a crystal structure of GroEL-GroES-ADP.AlF3 was determined at 2.8 A. A trigonal AlF3 metal complex was observed in the gamma-phosphate position of the nucleotide pocket of the cis ring. Surprisingly, when this structure was compared with that of the previously determined GroEL-GroES-ADP complex, no other differences were observed. We discuss the likely basis of the ability of gamma-phosphate binding to convert preformed GroEL-GroES-ADP-polypeptide complexes into the folding-active state.  相似文献   

19.
The crystal structure of Mycobacterium tuberculosis adenylate kinase (MtAK) in complex with two ADP molecules and Mg2+ has been determined at 1.9 A resolution. Comparison with the solution structure of the enzyme, obtained in the absence of substrates, shows significant conformational changes of the LID and NMP-binding domains upon substrate binding. The ternary complex represents the state of the enzyme at the start of the backward reaction (ATP synthesis). The structure is consistent with a direct nucleophilic attack of a terminal oxygen from the acceptor ADP molecule on the beta-phosphate from the donor substrate, and both the geometry and the distribution of positive charge in the active site support the hypothesis of an associative mechanism for phosphoryl transfer.  相似文献   

20.
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   

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