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.
Folding assistance and ATPase activity of GroEL are based on the existence of different conformations. In order to characterise these conformations, published data on steady state ATPase activity in the absence of GroES were reanalysed simultaneously in terms of the Nested MWC model. This model is a hierarchical extension of the symmetry-model of Monod et al. [J. Mol. Biol. 12 (1965) 88]. An unique set of GroEL specific parameters was obtained. This set was supported by comparison of predictions arising from this set of values with experimental data for hydrolysis of ATP in the presence of ADP and ATPgammaS, binding of ATPgammaS and ADP to GroEL in the absence of ATP, and binding of ATP as monitored by fluorescence labelling. Thus, for the first time, multiple data sets for the interaction of nucleotides with GroEL are described quantitatively by an allosteric model. A noteworthy feature of our model is that no negative cooperativity in ATP binding occurs in accordance to experimental observations. Furthermore, the model also includes the existence of a conformation with very high ATPase activity. Such a conformation might be of importance at a certain stage in the folding cycle.  相似文献   

2.
Biswas SB  Biswas-Fiss EE 《Biochemistry》2006,45(38):11505-11513
DnaB helicase is responsible for unwinding duplex DNA during chromosomal DNA replication and is an essential component of the DNA replication apparatus in Escherichia coli. We have analyzed the mechanism of binding of single-stranded DNA (ssDNA) by the DnaB x DnaC complex and DnaB helicase. Binding of ssDNA to DnaB helicase was significantly modulated by nucleotide cofactors, and the modulation was distinctly different for its complex with DnaC. DnaB helicase bound ssDNA with a high affinity [Kd = (5.09 +/- 0.32) x 10(-8) M] only in the presence of ATPgammaS, a nonhydrolyzable analogue of ATP, but not other nucleotides. The binding was sensitive to ionic strength but not to changes in temperature in the range of 30-37 degrees C. On the other hand, ssDNA binding in the presence of ADP was weaker than that observed with ATPgammaS, and the binding was insensitive to ionic strength. DnaC protein hexamerizes to form a 1:1 complex with the DnaB hexamer and loads it onto the ssDNA by forming a DnaB6 x DnaC6 dodecameric complex. Our results demonstrate that the DnaB6 x DnaC6 complex bound ssDNA with a high affinity [Kd = (6.26 +/- 0.65) x 10(-8) M] in the presence of ATP, unlike the DnaB hexamer. In the presence of ATPgammaS or ADP, binding of ssDNA by the DnaB6 x DnaC6 complex was a lower-affinity process. In summary, our results suggest that in the presence of ATP in vivo, the DnaB6 x DnaC6 complex should be more efficient in binding DNA as well as in loading DnaB onto the ssDNA than DnaB helicase itself.  相似文献   

3.
The Escherichia coli DEAD-box protein A (DbpA) is an RNA helicase that utilizes the energy from ATP binding and hydrolysis to facilitate structural rearrangements of rRNA. We have used the fluorescent nucleotide analogues, mantADP and mantATP, to measure the equilibrium binding affinity and kinetic mechanism of nucleotide binding to DbpA in the absence of RNA. Binding generates an enhancement in mant-nucleotide fluorescence and a corresponding reduction in intrinsic DbpA fluorescence, consistent with fluorescence resonance energy transfer (FRET) from DbpA tryptophan(s) to bound nucleotides. Fluorescent modification does not significantly interfere with the affinities and kinetics of nucleotide binding. Different energy transfer efficiencies between DbpA-mantATP and DbpA-mantADP complexes suggest that DbpA adopts nucleotide-dependent conformations. ADP binds (K(d) approximately 50 microM at 22 degrees C) 4-7 times more tightly than ATP (K(d) approximately 400 microM at 22 degrees C). Both nucleotides bind with relatively temperature-independent association rate constants (approximately 1-3 microM(-1) s(-1)) that are much lower than predicted for a diffusion-limited reaction. Differences in the binding affinities are dictated primarily by the dissociation rate constants. ADP binding occurs with a positive change in the heat capacity, presumably reflecting a nucleotide-induced conformational rearrangement of DbpA. At low temperatures (<22 degrees C), the binding free energies are dominated by favorable enthalpic and unfavorable entropic contributions. At physiological temperatures (>22 degrees C), ADP binding occurs with positive entropy changes. We favor a mechanism in which ADP binding increases the conformational flexibility and dynamics of DbpA.  相似文献   

4.
Xu H  Frank J  Niedenzu T  Saenger W 《Biochemistry》2000,39(40):12225-12233
The steady-state kinetic parameters of the ATPase activity of the homohexameric DNA helicase RepA and the binding of the fluorescent analogue epsilonADP to RepA have been studied. ssDNA stimulates RepA ATPase activity optimally at acidic pH 5.3-6.0. The sigmoidal kinetic curves in both the absence and presence of ssDNA show strong positive cooperativity for ATP hydrolysis, with oligonucleotides longer than 10mer optimal for ssDNA-stimulated ATPase activity. Fluorescence titrations show that, at 25 degrees C and in the absence of DNA, the binding of epsilonADP to RepA is biphasic with three high (K(1) = 1.54 x 10(6) M(-1)) and three low (K(2) = 4.71 x 10(4) M(-)(1)) affinity binding sites differing by 30-40-fold in binding constants. In the absence of cofactors, RepA melts cooperatively at T(m) = 65.8 +/- 0.1 degrees C and is more stable in the presence of ATPgammaS, T(m) = 68.1 +/- 0.2 degrees C (DeltaDeltaG 0.95 kcal/mol), than in the presence of ADP, T(m) = 66. 5 +/- 0.1 degrees C (DeltaDeltaG 0.29 kcal/mol), indicating that the additional phosphate group in ATPgammaS has a significant influence on RepA structure. A model is proposed in which individual subunits of RepA sequentially and cooperatively perform a multistep ATP hydrolytic cycle.  相似文献   

5.
Multidrug resistance-associated protein (MRP1) transports solutes in an ATP-dependent manner by utilizing its two nonequivalent nucleotide binding domains (NBDs) to bind and hydrolyze ATP. We found that ATP binding to the first NBD of MRP1 increases binding and trapping of ADP at the second domain (Hou, Y., Cui, L., Riordan, J. R., and Chang, X. (2002) J. Biol. Chem. 277, 5110-5119). These results were interpreted as indicating that the binding of ATP at NBD1 causes a conformational change in the molecule and increases the affinity for ATP at NBD2. However, we did not distinguish between the possibilities that the enhancement of ADP trapping might be caused by either ATP binding alone or hydrolysis. We now report the following. 1) ATP has a much lesser effect at 0 degrees C than at 37 degrees C. 2) After hexokinase treatment, the nonhydrolyzable ATP analogue, adenyl 5'-(yl iminodiphosphate), does not enhance ADP trapping. 3) Another nonhydrolyzable ATP analogue, adenosine 5'-(beta,gamma-methylene)triphosphate, whether hexokinase-treated or not, causes a slight enhancement. 4) In contrast, the hexokinase-treated poorly hydrolyzable ATP analogue, adenosine 5'-O-(thiotriphosphate) (ATPgammaS), enhances ADP trapping to a similar extent as ATP under conditions in which ATPgammaS should not be hydrolyzed. We conclude that: 1) ATP hydrolysis is not required to enhance ADP trapping by MRP1 protein; 2) with nucleotides having appropriate structure such as ATP or ATPgammaS, binding alone can enhance ADP trapping by MRP1; 3) the stimulatory effect on ADP trapping is greatly diminished when the MRP1 protein is in a "frozen state" (0 degrees C); and 4) the steric structure of the nucleotide gamma-phosphate is crucial in determining whether binding of the nucleotide to NBD1 of MRP1 protein can induce the conformational change that influences nucleotide trapping at NBD2.  相似文献   

6.
p97, an essential chaperone in endoplasmic reticulum-associated degradation and organelle biogenesis, contains two AAA domains (D1 and D2) and assembles as a stable hexamer. We present a quantitative analysis of nucleotide binding to both D1 and D2 domains of p97, the first detailed study of nucleotide binding to both AAA domains for this type of AAA+ ATPase. We report that adenosine 5'-O-(thiotriphosphate) (ATPgammaS) binds with similar affinity to D1 and D2, but ADP binds with higher affinity to D1 than D2, offering an explanation for the higher ATPase activity in D2. Stoichiometric measurements suggest that although both ADP and ATPgammaS can saturate all 6 nucleotide binding sites in D1, only 3-4 of the 6 D2 sites can bind ATPgammaS simultaneously. ATPgammaS binding triggers a downstream cooperative conformational change of at least three monomers, which involves conserved arginine fingers and is necessary for ATP hydrolysis.  相似文献   

7.
We have studied how nucleotides (ADP, AMP-PNP, and ATP) and the co-chaperonin GroES influence the GroEL-affected refolding of apo-alpha-lactalbumin. The refolding reactions induced by stopped-flow pH jumps were monitored by alpha-lactalbumin tryptophan fluorescence. The simple single-exponential character of the free-refolding kinetics of the protein allowed us to quantitatively analyze the kinetic traces of the GroEL-affected refolding with the aid of computer simulations, and to obtain the best-fit parameters for binding between GroEL and the refolding intermediate of alpha-lactalbumin by the non-linear least-squares method. When GroES was absent, the interaction between GroEL and alpha-lactalbumin could be well represented by a "cooperative-binding" model in which GroEL has two binding sites for alpha-lactalbumin with the affinity of the second site being tenfold weaker than that of the first, so that there is negative cooperativity between the two sites. The affinity between GroEL and alpha-lactalbumin was significantly reduced when ATP was present, while ADP and AMP-PNP did not alter the affinity. A comparison of this result with those reported previously for other target proteins suggests a remarkable adjustability of the GroEL 14-mer with respect to the nucleotide-induced reduction of affinity. When GroES was present, ATP as well as ADP and AMP-PNP were effective in reducing the affinity between GroEL and the refolding intermediate of alpha-lactalbumin. The affinity at a saturating concentration of ADP or AMP-PNP was about ten times lower than with GroEL alone. The ADP concentration at which the acceleration of the GroEL/ES-affected refolding of alphaLA was observed, was higher than the concentration at which the nucleotide-induced formation of the GroEL/ES complex took place. These results indicate that GroEL/ES complex formation itself is not enough to reduce the affinity for alpha-lactalbumin, and that further binding of the nucleotide to the GroEL/ES complex is required to reduce the affinity.  相似文献   

8.
RecA mediated homologous recombination requires cooperative ATP binding and hydrolysis to assume and maintain an active, extended DNA-protein (nucleoprotein) filament. Human RAD51 protein (hRAD51) lacks the magnitude of ATP-induced cooperativity and catalytic efficiency displayed by RecA. Here, we examined hRAD51 binding and ATPase inhibition pattern by ADP and ATP/adenosine 5'-O-(thiotriphosphate) (ATPgammaS). hRAD51 fully saturates with ATP/ATPgammaS regardless of DNA cofactor (K(D) approximately 5 microm; 1 ATP/1 hRAD51). The binding of ADP to hRAD51 appeared bimodal. The first mode was identical to ATP/ATPgammaS binding (K(app1) approximately 3 microm; 1 ADP/1 hRAD51), while a second mode occurred at elevated ADP concentrations (K(app2) > or = 125 microm; >1 ADP/1 hRAD51). We could detect ADP --> ATP exchange in the high affinity ADP binding mode (K(app1)) but not the low affinity binding mode (K(app2)). At low ATP concentrations (<0.3 mm), ADP and ATPgammaS competitively inhibit the hRAD51 ATPase (K(m)((app)) > K(m)). However, at high ATP (>0.3 mm), the hRAD51 ATPase was stimulated by concentrations of ATPgammaS that were 20-fold above the K(D). Ammonium sulfate plus spermidine decreased the affinity of hRAD51 for ADP substantially ( approximately 10-fold) and ATP modestly ( approximately 3-fold). Our results suggest that ATP binding is not rate-limiting but that the inability to sustain an active nucleoprotein filament probably restricts the hRAD51 ATPase.  相似文献   

9.
We have measured the energetics of ATP and ADP binding to single-headed actomyosin V and VI from the temperature dependence of the rate and equilibrium binding constants. Nucleotide binding to actomyosin V and VI can be modeled as two-step binding mechanisms involving the formation of collision complexes followed by isomerization to states with high nucleotide affinity. Formation of the actomyosin VI-ATP collision complex is much weaker and slower than for actomyosin V. A three-step binding mechanism where actomyosin VI isomerizes between two conformations, one competent to bind ATP and one not, followed by rapid ATP binding best accounts for the data. ADP binds to actomyosin V more tightly than actomyosin VI. At 25 degrees C, the strong ADP-binding equilibria are comparable for actomyosin V and VI, and the different overall ADP affinities arise from differences in the ADP collision complex affinity. The actomyosin-ADP isomerization leading to strong ADP binding is entropy driven at >15 degrees C and occurs with a large, positive change in heat capacity (DeltaC(P) degrees ) for both actomyosin V and VI. Sucrose slows ADP binding and dissociation from actomyosin V and VI but not the overall equilibrium constants for strong ADP binding, indicating that solvent viscosity dampens ADP-dependent kinetic transitions, presumably a tail swing that occurs with ADP binding and release. We favor a mechanism where strong ADP binding increases the dynamics and flexibility of the actomyosin complex. The heat capacity (DeltaC(P) degrees ) and entropy (DeltaS degrees ) changes are greater for actomyosin VI than actomyosin V, suggesting different extents of ADP-induced structural rearrangement.  相似文献   

10.
EnvZ is a histidine protein kinase important for osmoregulation in bacteria. While structural data are available for this enzyme, the nucleotide binding pocket is not well characterized. The ATP binding domain (EnvZB) was expressed, and its ability to bind nucleotide derivatives was assessed using equilbrium and stopped-flow fluorescence spectroscopy. The fluorescence emission of the trinitrophenyl derivatives, TNP-ATP and TNP-ADP, increase upon binding to EnvZB. The fluorescence enhancements were quantitatively abolished in the presence of excess ADP, indicating that the fluorescent probes occupy the nucleotide binding pocket. Both TNP-ATP and TNP-ADP bind to EnvZB with high affinity (K(d) = 2-3 microM). The TNP moiety attached to the ribose ring does not impede access of the fluorescent nucleotide into the binding pocket. The association rate constant for TNP-ADP is 7 microM(-1) s(-1), a value consistent with those for natural nucleotides and the eucaryotic protein kinases. Using competition experiments, it was found that ATP and ADP bind 30- and 150-fold more poorly, respectively, than the corresponding TNP-derivatized forms. Surprisingly, the physiological metal Mg(2+) is not required for ADP binding and only enhances ATP affinity by 3-fold. Although portions of the nucleotide pocket are disordered, the recombinant enzyme is highly stable, unfolding only at temperatures in excess of 70 degrees C. The unusually high affinity of the TNP derivatives compared to the natural nucleotides suggests that hydrophobic substitutions on the ribose ring enforce an altered binding mode that may be exploited for drug design strategies.  相似文献   

11.
Mg2+ dependence of guanine nucleotide binding to tubulin   总被引:1,自引:0,他引:1  
The relationship between the concentration of Mg2+ and the binding of GDP and GTP to tubulin dimers was investigated by measuring the displacement of the nucleotide bound at the exchangeable site (E-site) by radiolabeled GDP and GTP. A wide range of concentrations of GTP, GDP, and Mg2+ was explored. In the near absence of Mg2+, the affinity of tubulin for GDP was found to be much greater than its affinity for GTP. In the presence of 1.0 mM Mg2+, however, its affinity for GDP was slightly less than for GTP. The results could be quantitatively described in terms of a small number of reversible equilibria. Equilibrium constants, pertaining to measurements at 0 degrees C, in 0.1 M piperazine-N,N'-bis(2-ethanesulfonic acid), 0.2 mM dithioerythritol, 2 mM EGTA, pH 6.9, were obtained by nonlinear least squares fitting of the data. When the association constant of tubulin for GDP uncomplexed with Mg2+ was taken to be 1.6 X 10(7) M-1, that for uncomplexed GTP was found to be no larger than 1.4 x 10(4) M-1, at least 1100-fold smaller. The association constant of tubulin for the GDP.Mg2+ complex was found to be 2.5-2.7 x 10(7) M-1, while that for the GTP.Mg2+ complex is 6.4-9.0 x 10(7) M-1.  相似文献   

12.
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   

13.
Fedosova NU  Champeil P  Esmann M 《Biochemistry》2003,42(12):3536-3543
Transient kinetic analysis of nucleotide binding to pig kidney Na,K-ATPase using a rapid filtration technique shows that the interaction between nucleotide and enzyme apparently follows simple first-order kinetics both for ATP in the absence of Mg(2+) and for ADP in the presence or absence of Mg(2+). Rapid filtration experiments with Na,K-ATPase membrane sheets may nevertheless suffer from a problem of accessibility for a fraction of the ATPase binding sites. Accordingly, we estimate from these data that for ATP binding in the absence of Mg(2+) and the presence of 35 mM Na(+) at pH 7.0 at 20 degrees C, the bimolecular binding rate constant k(on) is about 30 microM(-1) x s(-1) and the dissociation rate constant k(off) is about 8 s(-1). In the presence of 10 mM Mg(2+), the binding rate constant is the same as that in the absence of Mg(2+). For ADP or MgADP the binding rate constant is about 20 microM(-1) x s(-1) and the dissociation rate constant is about 12 s(-1). Results of rapid-mixing stopped-flow experiments with the fluorescent dye eosin are also consistent with a one-step mechanism of binding of eosin to the ATPase nucleotide site. The implication of these results is that nucleotide binding to Na,K-ATPase both in the absence and presence of Mg(2+) appears to be a single-step event, at least on the time scale accessible in these experiments.  相似文献   

14.
We have examined the permissible nucleotide occupancy states of human MutSalpha. The MSH2.MSH6 heterodimer binds 1 mol of ADP and 1 mol of adenosine 5'-O-(thiotriphosphate) (ATPgammaS), with a K(d) for each nucleotide of about 1 microm. Anisotropy measurements using BODIPY TR and BODIPY FL fluorescent derivatives of ADP and 5'-adenylyl-beta,gamma-imidodiphosphate (AMPPNP) also indicate an interaction stoichiometry of 1 mol of ADP and 1 mol of triphosphate analogue per MutSalpha heterodimer. Di- and triphosphate sites can be simultaneously occupied as judged by sequential filling of the two binding site classes with differentially radiolabeled ADP and ATPgammaS and by fluorescence resonance energy transfer between BODIPY TR- and BODIPY FL-labeled ADP and AMPPNP. ATP hydrolysis by MutSalpha is accompanied by a pre-steady-state burst of ADP formation, and analysis of MutSalpha-bound nucleotide during the first turnover has demonstrated the presence of both ADP and ATP. Simultaneous presence of ADP and a nonhydrolyzable ATP analogue modulates MutSalpha.heteroduplex interaction in a manner that is distinct from that observed in the presence of ADP or nonhydrolyzable triphosphate alone, and it is unlikely that this effect is due to the presence of a mixed population of binary complexes between MutSalpha and ADP or a triphosphate analogue. These findings imply that MutSalpha has two nucleotide binding sites with differential specificities for ADP and ATP and suggest that the ADP.MutSalpha.ATP ternary complex has an important role in mismatch repair.  相似文献   

15.
Nucleotide binding to RecA results in either the high-DNA affinity form (Adenosine 5'-triphosphate (ATP)-bound) or the more inactive protein conformation associated with a lower affinity for DNA (Adenosine 5'-diphosphate (ADP)-bound). Many of the key structural differences between the RecA-ATP and RecA-ADP bound forms have yet to be elucidated. We have used caged-nucleotides and difference FTIR in efforts to obtain a comprehensive understanding of the molecular changes induced by nucleotide binding to RecA. The photochemical release of nucleotides (ADP and ATP) from biologically inactive precursors was used to initiate nucleotide binding to RecA. Here we present ATP hydrolysis assays and fluorescence studies suggesting that the caged nucleotides do not interact with RecA before photochemical release. Furthermore, we now compare difference spectra obtained in H2O and D2O as our first attempt at identifying the origin of the vibrations influenced by nucleotide binding. The infrared data suggest that unique alpha-helical, beta structures, and side chain rearrangements are associated with the high- and low-DNA affinity forms of RecA. Difference spectra obtained over time isolate contributions arising from perturbations in the nucleotide phosphates and have provided further information about the protein structural changes involved in nucleotide binding and the allosteric regulation of RecA.  相似文献   

16.
SecA is the dissociatable nucleotide and preprotein binding subunit of the bacterial translocase. The thermodynamics of nucleotide binding to soluble SecA at nucleotide binding site I were determined by isothermal titration calorimetry. Binding of ADP and non-hydrolyzable ATPgammaS is enthalpy-driven (DeltaH(0) of -14.44 and -5.56 kcal/mol, respectively), but is accompanied by opposite entropic contributions (DeltaS(0) of -18.25 and 9.55 cal/mol/K, respectively). ADP binding results in a large change in the heat capacity of SecA (DeltaC(p)=-780 cal/mol/K). It is suggested that ADP binding promotes the interaction between the two thermodynamically discernible domains of SecA which is accompanied by a shielding of hydrophobic surface from solvent.  相似文献   

17.
Four tight nucleotide binding sites of chloroplast coupling factor 1.   总被引:1,自引:0,他引:1  
We have examined the properties of the four tight nucleotide binding sites of reductively activated chloroplast coupling factor 1. Tight sites are here defined as those which retain bound nucleotides after passage of the chloroplast coupling factor 1 through Sephadex gel filtration centrifuge columns. Two of the sites, here called sites 4 and 5, have not been characterized in detail before. Site 4 has properties similar to those of site 1. It binds to ADP, ATP, and adenylyl-beta,gamma-imidodiphosphate (AMP-PNP) tightly in the presence or absence of Mg2+. Bound ADP exchanges rapidly with medium ADP, but rapid exchange with ATP or AMP-PNP requires Mg2+. Site 4 may slowly hydrolyze bound ATP in the absence of medium nucleotides. Site 5 has properties similar to those of site 2. Tight binding of ATP and AMP-PNP requires Mg2+, but Mg29+)-ADP is not tightly bound. Site 5 does not hydrolyze bound ATP in the absence of medium nucleotides. Complete filling of all four tight nucleotide binding sites requires about one millimolar nucleotide, suggesting that low affinity binding sites are converted to tight binding via a nucleotide binding-induced conformational change.  相似文献   

18.
Qu Q  Russell PL  Sharom FJ 《Biochemistry》2003,42(4):1170-1177
Drug transport mediated by P-glycoprotein (Pgp) is driven by hydrolysis of ATP at the two cytosolic nucleotide binding domains. However, little is currently known concerning the stoichiometry of nucleotide binding and how both stoichiometry and binding affinity change during the catalytic cycle of the transporter. To address this issue, we used fluorescence techniques to measure both the number of nucleotides bound to P-glycoprotein during various stages of the catalytic cycle and the affinity of nucleotide binding. Results showed that resting state P-glycoprotein bound two molecules of the fluorescent nucleotide derivative, 2'(3')-O-(2,4,6-trinitrophenyl)adenosine 5'-triphosphate (TNP-ATP), whereas the vanadate-trapped transition state bound only one nucleotide molecule. Both resting and transition state P-glycoprotein showed similar affinity for TNP-ATP/TNP-ADP and unlabeled ATP/ADP. Following binding of various drugs, resting state P-glycoprotein displayed a higher affinity for nucleotides, up to 4-fold depending on the compound used. In contrast, the transition state showed substantially lower (up to 3-fold) nucleotide binding affinity when the drug binding site(s) is/are occupied. These results indicate that both nucleotide binding domains of P-glycoprotein are likely to be occupied with either ATP (or ADP) in the resting state and the transition state in the absence of transport substrates. Drugs alter the binding affinity to favor association of ATP with P-glycoprotein at the start of the catalytic cycle and release of ADP from the transition state following nucleotide hydrolysis.  相似文献   

19.
ClpX requires ATP to unfold protein substrates and translocate them into the proteolytic chamber of ClpP for degradation. The steady-state parameters for hydrolysis of ATP and ATPgammaS by ClpX were measured with different protein partners and the kinetics of degradation of ssrA-tagged substrates were determined with both nucleotides. ClpX hydrolyzed ATPgammaS to ADP and thiophosphate at a rate (6/min) significantly slower than ATP hydrolysis (140/min), but the hydrolysis of both nucleotides was increased by ssrA-tagged substrates and decreased by ClpP. K(M) and k(cat) for hydrolysis of ATP and ATPgammaS were linearly correlated over a 200-fold range, suggesting that protein partners largely affect k(cat) rather than nucleotide binding, indicating that most bound ATP leaves the enzyme by hydrolysis rather than dissociation, and placing an upper limit of approximately 15 micro M on K(D) for both nucleotides. Competition studies with ClpX and fluorescently labeled ADP gave inhibition constants for ATPgammaS ( approximately 2 micro M) and ADP ( approximately 3 micro M) under the reaction conditions used for steady-state kinetics. In the absence of Mg(2+), where hydrolysis does not occur, the inhibition constant for ATP ( approximately 55 micro M) was weaker but very similar to the value for ATPgammaS ( approximately 45 micro M). Compared with ATP, ATPgammaS supported slow but roughly comparable rates of ClpXP degradation for two Arc-ssrA substrates and denatured GFP-ssrA, but not of native GFP-ssrA. These results show that the processing of protein substrates by ClpX is closely coupled to the maximum rate of nucleotide hydrolysis.  相似文献   

20.
Mismatch repair (MMR) is essential for eliminating biosynthetic errors generated during replication or genetic recombination in virtually all organisms. The critical first step in Escherichia coli MMR is the specific recognition and binding of MutS to a heteroduplex, containing either a mismatch or an insertion/deletion loop of up to four nucleotides. All known MutS homologs recognize a similar broad spectrum of substrates. Binding and hydrolysis of nucleotide cofactors by the MutS-heteroduplex complex are required for downstream MMR activity, although the exact role of the nucleotide cofactors is less clear. Here, we showed that MutS bound to a 30-bp heteroduplex containing an unpaired T with a binding affinity ≈ 400-fold stronger than to a 30-bp homoduplex, a much higher specificity than previously reported. The binding of nucleotide cofactors decreased both MutS specific and nonspecific binding affinity, with the latter marked by a larger drop, further increasing MutS specificity by ≈ 3-fold. Kinetic studies showed that the difference in MutS Kd for various heteroduplexes was attributable to the difference in intrinsic dissociation rate of a particular MutS-heteroduplex complex. Furthermore, the kinetic association event of MutS binding to heteroduplexes was marked by positive cooperativity. Our studies showed that the positive cooperativity in MutS binding was modulated by the binding of nucleotide cofactors. The binding of nucleotide cofactors transformed E. coli MutS tetramers, the functional unit in E. coli MMR, from a cooperative to a noncooperative binding form. Finally, we found that E. coli MutS bound to single-strand DNA with significant affinity, which could have important implication for strand discrimination in eukaryotic MMR mechanism.  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