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.
In crystal structures of bovine MF(1), the side chains of alpha F(357) and beta R(372) are near the adenines of nucleotides bound to noncatalytic sites. To determine if during catalysis these side chains must pass through the different arrangements in which they are present in crystal structures, the catalytic properties of the (alpha F(357)C)(3)(beta R(372)C)(3)gamma subcomplex of the TF(1)-ATPase were characterized before and after cross-linking the introduced cysteines with CuCl(2). The unmodified mutant enzyme hydrolyzes MgATP at 50% the rate exhibited by wild type. Detailed comparison of the catalytic properties of the double mutant enzyme before and after cross-linking with those of the wild-type subcomplex revealed the following. Before cross-linking, the (alpha F(357)C)(3)(beta R(372)C)(3)gamma subcomplex has less tendency than wild type to release inhibitory MgADP entrapped in a catalytic site during turnover when MgATP binds to noncatalytic sites. Following cross-linking, ATPase activity is reduced 5-fold, and inhibitory MgADP entrapped in a catalytic site during turnover does not release under conditions wherein binding of ATP to noncatalytic sites of the wild-type enzyme promotes release of MgADP from the affected catalytic site. When assayed in the presence of lauryldimethylamine oxide, which prevents turnover-dependent entrapment of inhibitory MgADP in a catalytic site, ATPase activity of the cross-linked form is 47% that of the unmodified mutant enzyme. These results suggest that, during catalysis, the side chains of alpha F(357) and beta R(372) do not pass through the extremely different relative positions in which they exist at the three noncatalytic site interfaces in crystal structures.  相似文献   

2.
In the crystal structure of the bovine heart mitochondrial F(1)-ATPase (Abrahams, J. P., Leslie, A. G. W., Lutter, R., and Walker, J. E. (1994) Nature 370, 621-628), the two liganded beta subunits, one with MgAMP-PNP bound to the catalytic site (beta(T)) and the other with MgADP bound (beta(D)) have closed conformations. The empty beta subunit (beta(E)) has an open conformation. In beta(T) and beta(D), the distance between the carboxylate of beta-Asp(315) and the guanidinium of beta-Arg(337) is 3.0-4.0 A. These side chains are at least 10 A apart in beta(E). The alpha(3)(betaD311C/R333C)(3)gamma subcomplex of TF(1) with the corresponding residues substituted with cysteine has very low ATPase activity unless it is reduced prior to assay or assayed in the presence of dithiothreitol. The reduced subcomplex hydrolyzes ATP at 50% the rate of wild-type and is rapidly inactivated by oxidation by CuCl(2) with or without magnesium nucleotides bound to catalytic sites. Titration of the subcomplex with iodo[(14)C]acetamide after prolonged treatment with CuCl(2) in the presence or absence of 1 mM MgADP revealed nearly two free sulfhydryl groups/mol of enzyme. Therefore, one pair of introduced cysteines is located on a beta subunit that exists in the open or partially open conformation even when catalytic sites are saturated with MgADP. Since V(max) of ATP hydrolysis is attained when three catalytic sites of F(1) are saturated, the catalytic site that binds ATP must be closing as the catalytic site that releases products is opening.  相似文献   

3.
Boltz KW  Frasch WD 《Biochemistry》2006,45(37):11190-11199
F(1)-ATPase mutations in Escherichia coli that changed the strength of hydrogen bonds between the alpha and beta subunits in a location that links the catalytic site to the interface between the beta catch loop and the gamma subunit were examined. Loss of the ability to form the hydrogen bonds involving alphaS337, betaD301, and alphaD335 lowered the k(cat) of ATPase and decreased its susceptibility to Mg(2+)-ADP-AlF(n) inhibition, while mutations that maintain or strengthen these bonds increased the susceptibility to Mg(2+)-ADP-AlF(n) inhibition and lowered the k(cat) of ATPase. These data suggest that hydrogen bonds connecting alphaS337 to betaD301 and betaR323 and connecting alphaD335 to alphaS337 are important to transition state stabilization and catalytic function that may result from the proper alignment of catalytic site residues betaR182 and alphaR376 through the VISIT sequence (alpha344-348). Mutations betaD301E, betaR323K, and alphaR282Q changed the rate-limiting step of the reaction as determined by an isokinetic plot. Hydrophobic mutations of betaR323 decreased the susceptibility to Mg(2+)-ADP-AlF(n)() inhibition and lowered the number of interactions required in the rate-limiting step yet did not affect the k(cat) of ATPase, suggesting that betaR323 is important to transition state formation. The decreased rate of ATP synthase-dependent growth and decreased level of lactate-dependent quenching observed with alphaD335, betaD301, and alphaE283 mutations suggest that these residues may be important to the formation of an alternative set of hydrogen bonds at the interface of the alpha and beta subunits that permits the release of intersubunit bonds upon the binding of ATP, allowing gamma rotation in the escapement mechanism.  相似文献   

4.
In the crystal structure of bovine mitochondrial F(1)-ATPase (MF(1)) (Abrahams, J. P., Leslie, A. G. W., Lutter, R., and Walker, J. E. (1994) Nature 370, 621-628), the side chain oxygen of betaThr(163) interacts directly with Mg(2+) coordinated to 5'-adenylyl beta, gamma-imidodiphosphate or ADP bound to catalytic sites of beta subunits present in closed conformations. In the unliganded beta subunit present in an open conformation, the hydroxyl of betaThr(163) is hydrogen-bonded to the carboxylate of betaGlu(199). Substitution of betaGlu(201) (equivalent to betaGlu(199) in MF(1)) in the alpha(3)beta(3)gamma subcomplex of the F(1)-ATPase from the thermophilic Bacillus PS3 with cysteine or valine increases the propensity to entrap inhibitory MgADP in a catalytic site during hydrolysis of 50 microM ATP. These substitutions lower K(m3) (the Michaelis constant for trisite ATP hydrolysis) relative to that of the wild type by 25- and 10-fold, respectively. Fluorescence quenching of alpha(3)(betaE201C/Y341W)(3)gamma and alpha(3)(betaY341W)(3)gamma mutant subcomplexes showed that MgATP and MgADP bind to the third catalytic site of the double mutant with 8.4- and 4.4-fold higher affinity, respectively, than to the single mutant. These comparisons support the hypothesis that the hydrogen bond observed between the side chains of betaThr(163) and betaGlu(199) in the unliganded catalytic site in the crystal structure of MF(1) stabilizes the open conformation of the catalytic site during ATP hydrolysis.  相似文献   

5.
S Nadanaciva  J Weber  A E Senior 《Biochemistry》1999,38(24):7670-7677
Beta-Arg-182 in Escherichia coli F1-ATPase (beta-Arg-189 in bovine mitochondrial F1) is a residue which lies close to catalytic site bound nucleotide (Abrahams et al. (1994) Nature 370, 621-628). Here we investigated the role of this residue by characterizing two mutants, betaR182Q and betaR182K. Oxidative phosphorylation and steady-state ATPase activity of purified F1 were severely impaired by both mutations. Catalytic site nucleotide-binding parameters were measured using the fluorescence quench of beta-Trp-331 that occurred upon nucleotide binding to purified F1 from betaR182Q/betaY331W and betaR182K/betaY331W double mutants. It was found that (a) beta-Arg-182 interacts with the gamma-phosphate of MgATP, particularly at catalytic sites 1 and 2, (b) beta-Arg-182 has no functional interaction with the beta-phosphate of MgADP or with the magnesium of the magnesium-nucleotide complex in the catalytic sites, and (c) beta-Arg-182 is directly involved in the stabilization of the catalytic transition state. In these features the role of beta-Arg-182 resembles that of another positively charged residue in the catalytic site, the conserved lysine of the Walker A motif, beta-Lys-155. A further role of beta-Arg-182 is suggested, namely involvement in conformational change at the catalytic site beta-alpha subunit interface that is required for multisite catalysis.  相似文献   

6.
Ren H  Bandyopadhyay S  Allison WS 《Biochemistry》2006,45(19):6222-6230
The alpha(3)(betaM(222)S/Y(345)W)(3)gamma double-mutant subcomplex of the F(1)-ATPase from the thermophilic Bacillus PS3 (TF(1)), free of endogenous nucleotides, does not entrap inhibitory MgADP in a catalytic site during turnover. It hydrolyzes 100 nM-2 mM ATP with a K(m) of 31 microM and a k(cat) of 220 s(-)(1). Fluorescence titrations of the introduced tryptophans with MgADP or MgATP revealed that both Mg-nucleotide complexes bind to the catalytic site of the highest affinity with K(d)()1 values of less than 1 nM and bind to the site of intermediate affinity with a common K(d)2 value of about 12 nM. The K(d)3 values obtained for the catalytic site of the lowest affinity from titrations with MgADP and MgATP are 25 and 37 microM, respectively. The double mutant hydrolyzes 200 nM ATP with a first-order rate of 1.5 s(-)(1), which is 0.7% of k(cat). Hence, it does not hydrolyze ATP at a significant rate when the catalytic site of intermediate affinity is saturated and the catalytic site of the lowest affinity is minimally occupied. After the addition of stoichiometric MgATP to the alpha(3)(betaM(222)S/Y(345)W)(3)gamma subcomplex, one-third of the tryptophan fluorescence remains quenched after 10 min. The product [(3)H]ADP remains bound when the wild-type and double-mutant subcomplexes hydrolyze substoichiometric [(3)H]ATP. In contrast, (32)P(i) is not retained when the wild-type subcomplex hydrolyzes substoichiometric [gamma-(32)P]ATP. This precludes assessment of the equilibrium at the high-affinity catalytic site when the wild-type TF(1) subcomplex hydrolyzes substoichiometric ATP.  相似文献   

7.
Refolding together the expressed alpha and beta subunits of the Rhodospirillum rubrum F(1)(RF(1))-ATPase led to assembly of only alpha(1)beta(1) dimers, showing a stable low MgATPase activity. When incubated in the presence of AlCl(3), NaF and either MgAD(T)P or CaAD(T)P, all dimers associated into closed alpha(3)beta(3) hexamers, which also gained a low CaATPase activity. Both hexamer ATPase activities exhibited identical rates and properties to the open dimer MgATPase. These results indicate that: a) the hexamer, as the dimer, has no catalytic cooperativity; b) aluminium fluoride does not inhibit their MgATPase activity; and c) it does enable the assembly of RrF(1)-alpha(3)beta(3) hexamers by stabilizing their noncatalytic alpha/beta interfaces. Refolding of the RrF(1)-alpha and beta subunits together with the spinach chloroplast F(1) (CF(1))-gamma enabled a simple one-step assembly of two different hybrid RrF(1)-alpha(3)beta(3)/CF(1)gamma complexes, containing either wild type RrF(1)-beta or the catalytic site mutant RrF(1)beta-T159S. They exhibited over 100-fold higher CaATPase and MgATPase activities than the stabilized hexamers and showed very different catalytic properties. The hybrid wild type MgATPase activity was, as that of RrF(1) and CF(1) and unlike its higher CaATPase activity, regulated by excess free Mg(2+) ions, stimulated by sulfite, and inhibited by azide. The hybrid mutant had on the other hand a low CaATPase but an exceptionally high MgATPase activity, which was much less sensitive to the specific MgATPase effectors. All these very different ATPase activities were regulated by thiol modulation of the hybrid unique CF(1)-gamma disulfide bond. These hybrid complexes can provide information on the as yet unknown factors that couple ATP binding and hydrolysis to both thiol modulation and rotational motion of their CF(1)-gamma subunit.  相似文献   

8.
Fluorescence titrations of the alpha(3)(betaG(156)C/Y(345)W)(3)gamma, alpha(3)(betaE(199)V/Y(345)W)(3)gamma, and alpha(3)(betaY(345)W)(3)gamma subcomplexes of TF(1) with nucleotides show that the betaG(156)C substitution substantially lowers the affinity of catalytic sites for ATP and ADP with or without Mg(2+), whereas the betaE(199)V substitution increases the affinity of catalytic sites for nucleotides. Whereas the alpha(3)(betaG(156)C)(3)gamma and alpha(3)(betaE(199)V)(3)gamma subcomplexes hydrolyze 2 mM ATP at 2% and 0.7%, respectively, of the rate exhibited by the wild-type enzyme, the alpha(3)(betaG(156)C/E(199)V)(3)gamma hydrolyzes 2 mM ATP at 9% the rate exhibited by the wild-type enzyme. The alpha(3)(betaG(156)C)(3)gamma, alpha(3)(betaG(156)C/E(199)V)(3)gamma, and alpha(3)(betaG(156)C/E(199)V/Y(345)W)(3)gamma subcomplexes resist entrapment of inhibitory MgADP in a catalytic site during turnover. Product [(3)H]ADP remains tightly bound to a single catalytic site when the wild-type, betaE(199)V, betaY(345)W, and betaE(199)V/Y(345)W subcomplexes hydrolyze substoichiometric [(3)H]ATP, whereas it is not retained by the betaG(156)C and betaG(156)C/Y(345)W subcomplexes. Less firmly bound, product [(3)H]ADP is retained when the betaG(156)C/E(199)V and betaG(156)C/E(199)V/Y(345)W mutants hydrolyze substoichiometric [(3)H]ATP. The Lineweaver-Burk plot obtained with the betaG(156)C mutant is curved downward in a manner indicating that its catalytic sites act independently during ATP hydrolysis. In contrast, the betaG(156)C/E(199)V and betaG(156)C/E(199)V/Y(345)W mutants hydrolyze ATP with linear Lineweaver-Burk plots, indicating cooperative trisite catalysis. It appears that the betaG(156)C substitution destabilizes the closed conformation of a catalytic site hydrolyzing MgATP in a manner that allows release of products in the absence of catalytic site cooperativity. Insertion of the betaE(199)V substitution into the betaG(156)C mutant restores cooperativity by restricting opening of the catalytic site hydrolyzing MgATP for product release until an open catalytic site binds MgATP.  相似文献   

9.
It has been reported that shifts in the fluorescence emission spectrum of the introduced tryptophans in the betaF155W mutant of Escherichia coli F(1) (bovine heart mitochondria F(1) residue number) can quantitatively distinguish between the number of catalytic sites occupied with ADP and ATP during steady-state ATP hydrolysis (Weber, J., Bowman, C., and Senior, A. E. (1996) J. Biol. Chem. 271, 18711--18718). In contrast, addition of MgADP, Mg-5'-adenylyl beta,gamma-imidophosphate (MgAMP-PNP), and MgATP in 1:1 ratios to the alpha(3)(betaF155W)(3)gamma subcomplex of thermophilic Bacillus PS3 F(1) (TF(1)) induced nearly identical blue shifts in the fluorescence emission maximum that was accompanied by quenching. Addition of 2 mm MgADP induced a slightly greater blue shift and a slight increase in intensity over those observed with 1:1 MgADP. However, addition of 2 mm MgAMP-PNP or MgATP induced a much greater blue shift and substantially enhanced fluorescence intensity over those observed in the presence of stoichiometric MgADP or MgAMP-PNP. It is clear from these results that the fluorescence spectrum of the introduced tryptophans in the betaF155W mutant of TF(1) does not respond in regular increments at any wavelength as catalytic sites are filled with nucleotides. The fluorescence spectrum observed after entrapping MgADP-fluoroaluminate complexes in two catalytic sites of the betaF155W subcomplex indicates that the fluorescence emission spectrum of the enzyme is maximally perturbed when nucleotides are bound to two catalytic sites. This finding is consistent with accumulating evidence suggesting that only two beta subunits in the alpha(3)beta(3)gamma subcomplex of TF(1) can simultaneously exist in the completely closed conformation.  相似文献   

10.
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   

11.
F1-ATPase, the catalytic sector of Fo-F1 ATPases-ATPsynthases, displays an apparent negative cooperativity for ATP hydrolysis at high ATP concentrations which involves noncatalytic and catalytic nucleotide binding sites. The molecular mechanism of such cooperativity is currently unknown. To get further insights, we have investigated the structural consequences of the single mutation of two residues: Q173L in the alpha-subunit and Q170Y in the beta-subunit of the F1-ATPase of the yeast Schizosaccharomyces pombe. These residues are localized in or near the Walker-A motifs of each subunit and their mutation produces an opposite effect on the negative cooperativity. The betaQ170 residue (M167 in beef heart) is located close to the binding site for the phosphate-Mg moiety of the nucleotide. Its replacement by tyrosine converts this site into a close state with increased affinity for the bound nucleotide and leads to an increase of negative cooperativity. In contrast, the alphaQ173L mutation (Q172 in beef heart) abolishes negative cooperativity due to the loss of two H-bonds: one stabilizing the nucleotide bound to the noncatalytic site and the other linking alphaQ173 to the adjacent betaT354, localized at the alpha(DP)-beta(TP) interface. The properties of these mutants suggest that negative cooperativity occurs through interactions between neighbor alpha- and beta-subunits. Indeed, in the beef heart enzyme, (i) the alpha(DP)-beta(TP) interface is stabilized by a vicinal alphaR171-betaD352 salt bridge (ii) betaD352 and betaT354 belong to a short peptidic stretch close to betaY345, the aromatic group of which interacts with the adenine moiety of the nucleotide bound to the catalytic site. We therefore propose that the betaY345-betaT354 stretch (beef heart numbering) constitutes a short link that drives structural modifications from a noncatalytic site to the neighbor catalytic site in which, as a result, the affinity for ADP is modulated.  相似文献   

12.
F1-ATPase is inactivated by entrapment of MgADP in catalytic sites and reactivated by MgATP or P(i). Here, using a mutant alpha(3)beta(3)gamma complex of thermophilic F(1)-ATPase (alpha W463F/beta Y341W) and monitoring nucleotide binding by fluorescence quenching of an introduced tryptophan, we found that P(i) interfered with the binding of MgATP to F(1)-ATPase, but binding of MgADP was interfered with to a lesser extent. Hydrolysis of MgATP by F(1)-ATPase during the experiments did not obscure the interpretation because another mutant, which was able to bind nucleotide but not hydrolyse ATP (alpha W463F/beta E190Q/beta Y341W), also gave the same results. The half-maximal concentrations of P(i) that suppressed the MgADP-inhibited form and interfered with MgATP binding were both approximately 20 mm. It is likely that the presence of P(i) at a catalytic site shifts the equilibrium from the MgADP-inhibited form to the enzyme-MgADP-P(i) complex, an active intermediate in the catalytic cycle.  相似文献   

13.
Residues responsible for phosphate binding in F(1)F(0)-ATP synthase catalytic sites are of significant interest because phosphate binding is believed linked to proton gradient-driven subunit rotation. From x-ray structures, a phosphate-binding subdomain is evident in catalytic sites, with conserved betaArg-246 in a suitable position to bind phosphate. Mutations betaR246Q, betaR246K, and betaR246A in Escherichia coli were found to impair oxidative phosphorylation and to reduce ATPase activity of purified F(1) by 100-fold. In contrast to wild type, ATPase of mutants was not inhibited by MgADP-fluoroaluminate or MgADP-fluoroscandium, showing the Arg side chain is required for wild-type transition state formation. Whereas 7-chloro-4-nitrobenzo-2-oxa-1,3-diazole (NBD-Cl) inhibited wild-type ATPase essentially completely, ATPase in mutants was inhibited maximally by approximately 50%, although reaction still occurred at residue betaTyr-297, proximal to betaArg-246 in the phosphate-binding pocket. Inhibition characteristics supported the conclusion that NBD-Cl reacts in betaE (empty) catalytic sites, as shown previously by x-ray structure analysis. Phosphate protected against NBD-Cl inhibition in wild type but not in mutants. The results show that phosphate can bind in the betaE catalytic site of E. coli F(1) and that betaArg-246 is an important phosphate-binding residue.  相似文献   

14.
Only beta-beta cross-links form when the alpha(3)(betaE(395)C)(3)gammaK(36)C (MF(1) residue numbers) double mutant subcomplex of TF(1), the F(1)-ATPase from the thermophilic Bacillus PS3, is slowly inactivated with CuCl(2) in the presence or absence of MgATP. The same slow rate of inactivation and extent of beta-beta cross-linking occur upon treatment of the alpha(3)(betaE(395)C)(3)gamma single mutant subcomplex with CuCl(2) under the same conditions. In contrast, the alpha(3)(betaE(395)C)(3)gammaR(33)C and alpha(3)(betaE(395)C)(3)gammaR(75)C double mutant subcomplexes of TF(1) are rapidly inactivated by CuCl(2) under the same conditions that is accompanied by complete beta-gamma cross-linking. The ATPase activity of each mutant enzyme containing the betaE(395)C substitution is stimulated to a much greater extent by the nonionic detergent lauryldimethylamine oxide (LDAO) than wild-type enzyme, whereas the ATPase activities of the gammaR(33)C, gammaK(36)C, and gammaR(75)C single mutants are stimulated to about the same extent as wild-type enzyme by LDAO. This indicates that the E(395)C substitution in the (394)DELSEED(400) segment of beta subunits increases propensity of the enzyme to entrap inhibitory MgADP in a catalytic site during turnover. These results are discussed in perspective with (i) the ionic track predicted from molecular dynamics simulations to operate during energy-driven ATP synthesis by MF(1), the F(1)-ATPase from bovine heart mitochondria [Ma, J., Flynn, T. C., Cui, Q., Leslie, A. G. W., Walker, J. E., and Karplus, M. (2002) Structure 10, 921-931]; and (ii) the possibility that the betaE(395)C substitution might induce a global effect that alters affinity of noncatalytic sites for nucleotides or alters communication between noncatalytic sites and catalytic sites during ATP hydrolysis.  相似文献   

15.
Nadanaciva S  Weber J  Senior AE 《Biochemistry》2000,39(31):9583-9590
MgADP in combination with fluoroscandium (ScFx) is shown to form a potently inhibitory, tightly bound, noncovalent complex at the catalytic sites of F(1)-ATPase. The F(1).MgADP.ScFx complex mimics a catalytic transition state. Notably, ScFx caused large enhancement of MgADP binding affinity at both catalytic sites 1 and 2, with little effect at site 3. These results indicate that sites 1 and 2 may form a transition state conformation. A new direct optical probe of F(1)-ATPase catalytic transition state conformation is also reported, namely, substantial enhancement of fluorescence emission of residue beta-Trp-148 observed upon binding of MgADP.ScFx or MgIDP. ScFx. Using this fluorescence signal, titrations were performed with MgIDP.ScFx which demonstrated that catalytic sites 1 and 2 can both form a transition state conformation but site 3 cannot. Supporting data were obtained using MgIDP-fluoroaluminate. Current models of the MgATP hydrolysis mechanism uniformly make the assumption that only one catalytic site hydrolyzes MgATP at any one time. The fluorometal analogues demonstrate that two sites have the capability to form the transition state simultaneously.  相似文献   

16.
The mechanism of inhibition of yeast mitochondrial F(1)-ATPase by its natural regulatory peptide, IF1, was investigated by correlating the rate of inhibition by IF1 with the nucleotide occupancy of the catalytic sites. Nucleotide occupancy of the catalytic sites was probed by fluorescence quenching of a tryptophan, which was engineered in the catalytic site (beta-Y345W). Fluorescence quenching of a beta-Trp(345) indicates that the binding of MgADP to F(1) can be described as 3 binding sites with dissociation constants of K(d)(1) = 10 +/- 2 nm, K(d2) = 0.22 +/- 0.03 microm, and K(d3) = 16.3 +/- 0.2 microm. In addition, the ATPase activity of the beta-Trp(345) enzyme followed simple Michaelis-Menten kinetics with a corresponding K(m) of 55 microm. Values for the K(d) for MgATP were estimated and indicate that the K(m) (55 microm) for ATP hydrolysis corresponds to filling the third catalytic site on F(1). IF1 binds very slowly to F(1)-ATPase depleted of nucleotides and under unisite conditions. The rate of inhibition by IF1 increased with increasing concentration of MgATP to about 50 mum, but decreased thereafter. The rate of inhibition was half-maximal at 5 microm MgATP, which is 10-fold lower than the K(m) for ATPase. The variations of the rate of IF1 binding are related to changes in the conformation of the IF1 binding site during the catalytic reaction cycle of ATP hydrolysis. A model is proposed that suggests that IF1 binds rapidly, but loosely to F(1) with two or three catalytic sites filled, and is then locked in the enzyme during catalytic hydrolysis of ATP.  相似文献   

17.
The crystal structure of yeast mitochondrial F(1) ATPase contains three independent copies of the complex, two of which have similar conformations while the third differs in the position of the central stalk relative to the alpha(3)beta(3) sub-assembly. All three copies display very similar asymmetric features to those observed for the bovine enzyme, but the yeast F(1) ATPase structures provide novel information. In particular, the active site that binds ADP in bovine F(1) ATPase has an ATP analog bound and therefore this structure does not represent the ADP-inhibited form. In addition, one of the complexes binds phosphate in the nucleotide-free catalytic site, and comparison with other structures provides a picture of the movement of the phosphate group during initial binding and subsequent catalysis. The shifts in position of the central stalk between two of the three copies of yeast F(1) ATPase and when these structures are compared to those of the bovine enzyme give new insight into the conformational changes that take place during rotational catalysis.  相似文献   

18.
The crystal structures of concanavalin A in complex with Man(alpha1-6)Man(alpha1-O)Me and Man(alpha1-3)Man(alpha1-O)Me were determined at resolutions of 2.0 and 2.8 A, respectively. In both structures, the O-1-linked mannose binds in the conserved monosaccharide-binding site. The O-3-linked mannose of Man(alpha1-3)Man(alpha1-O)Me binds in the hydrophobic subsite formed by Tyr-12, Tyr-100, and Leu-99. The shielding of a hydrophobic surface is consistent with the associated large heat capacity change. The O-6-linked mannose of Man(alpha1-6)Man(alpha1-O)Me binds in the same subsite formed by Tyr-12 and Asp-16 as the reducing mannose of the highly specific trimannose Man(alpha1-3)[Man(alpha1-6)]Man(alpha1-O)Me. However, it is much less tightly bound. Its O-2 hydroxyl makes no hydrogen bond with the conserved water 1. Water 1 is present in all the sugar-containing concanavalin A structures and increases the complementarity between the protein-binding surface and the sugar, but is not necessarily a hydrogen-bonding partner. A water analysis of the carbohydrate-binding site revealed a conserved water molecule replacing O-4 on the alpha1-3-linked arm of the trimannose. No such water is found for the reducing or O-6-linked mannose. Our data indicate that the central mannose of Man(alpha1-3)[Man(alpha1-6)]Man(alpha1-O)Me primarily functions as a hinge between the two outer subsites.  相似文献   

19.
Chen W  Hu CY  Crampton DJ  Frasch WD 《Biochemistry》2000,39(31):9393-9400
Metal ligands of the VO(2+)-adenosine diphosphate (ADP) complex bound to high-affinity catalytic site 1 of chloroplast F(1) adenosine triphosphatase (CF(1) ATPase) were characterized by electron paramagnetic resonance (EPR) spectroscopy. This EPR spectrum contains two EPR species designated E and F not observed when VO(2+)-nucleotide is bound to site 3 of CF(1). Site-directed mutations betaE197C, betaE197D, and betaE197S in Chlamydomonas CF(1) impair ATP synthase and ATPase activity catalyzed by CF(1)F(o) and soluble CF(1), respectively, indicating that this residue is important for enzyme function. These mutations caused large changes in the (51)V hyperfine tensors of VO(2+)-nucleotide bound to site 1 but not to site 3. Mutations to the Walker homology B aspartate betaD262C, betaD262H, and betaD262T of Chlamydomonas CF(1) caused similar effects on the EPR spectrum of VO(2+)-ADP bound to site 1. These results indicate that the conversion of the low-affinity site 3 conformation to high-affinity site 1 involves the incorporation betaE197 and betaD262 as metal ligands.  相似文献   

20.
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   

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