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 78 毫秒
1.
Vacuolar-type H(+)-ATPase (V-ATPase) catalyzes ATP synthesis and hydrolysis coupled with proton translocation across membranes via a rotary motor mechanism. Here we report biochemical and biophysical catalytic properties of V-ATPase from Thermus thermophilus. ATP hydrolysis of V-ATPase was severely inhibited by entrapment of Mg-ADP in the catalytic site. In contrast, the enzyme was very active for ATP synthesis (approximately 70 s(-1)) with the K(m) values for ADP and phosphate being 4.7 +/- 0.5 and 460 +/- 30 microm, respectively. Single molecule observation showed V-ATPase rotated in a 120 degrees stepwise manner, and analysis of dwelling time allowed the binding rate constant k(on) for ATP to be estimated ( approximately 1.1 x 10(6) m(-1) s(-1)), which was much lower than the k(on) (= V(max)/K(m)) for ADP ( approximately 1.4 x 10(7) m(-1) s(-1)). The slower k(on)(ATP) than k(on)(ADP) and strong Mg-ADP inhibition may contribute to prevent wasteful consumption of ATP under in vivo conditions when the proton motive force collapses.  相似文献   

2.
Chesnokova LS  Witt SN 《Biochemistry》2005,44(33):11224-11233
Hsp70 chaperones are heterotropic allosteric systems in which ATP and misfolded or aggregated polypeptides are the activating ligands. To gain insight into the mechanism by which ATP and polypeptides regulate Hsp70 chaperone activity, the effect of a short peptide on the K(M) for ATP was analyzed using the Escherichia coli Hsp70 called DnaK. In the absence of peptide, the K(-P)(M) for ATP is 52 +/- 11 nM, whereas this value jumps to 14.6 +/- 1.6 microM in the presence of saturating peptide. This finding supports a mechanism in which ATP binding drives the chaperone in one direction and peptide binding pushes the chaperone back in the opposite direction (and thus increases K(M)), according to ATP + DnaK.P <==> ATP.DnaK.P <==> ATP.DnaK* + P, where ATP.DnaK.P is an intermediate from which competing ATP hydrolysis occurs (ATP.DnaK.P --> ADP.DnaK.P). We show that this branched mechanism can even explain how DnaK hydrolyzes ATP in the absence of peptide and that the true rate constant for DnaK-mediated ATP hydrolysis (k(hy)) in the absence of peptide may be as high as 0.5 s(-)(1) (rather than 5 x 10(-)(4) s(-)(1) as often stated in the literature). What happens is that a conformational equilibrium outcompetes ATP hydrolysis and effectively reduces the concentration of the intermediate by a factor of a thousand, resulting in the following relation: k(cat) = k(hy)/1000 = 5 x 10(-)(4) s(-)(1). How polypeptide substrates and the co-chaperone DnaJ modulate DnaK to achieve its theoretical maximal rate of ATP hydrolysis, which we suggest is 0.5 s(-)(1), is discussed.  相似文献   

3.
Foster KA  Gilbert SP 《Biochemistry》2000,39(7):1784-1791
Ncd is a kinesin-related motor protein which drives movement to the minus-end of microtubules. The kinetics of Ncd were investigated using the dimeric construct MC1 (Leu(209)-Lys(700)) expressed in Escherichia coli strain BL21(DE) as a nonfusion protein [Chandra, R., Salmon, E. D., Erickson, H. P., Lockhart, A., and Endow, S. A. (1993) J. Biol. Chem. 268, 9005-9013]. Acid chemical quench flow methods were used to measure directly the rate of ATP hydrolysis, and stopped-flow kinetic methods were used to determine the kinetics of mantATP binding, mantADP release, dissociation of MC1 from the microtubule, and binding of MC1 to the microtubule. The results define a minimal kinetic mechanism, M.N + ATP M.N.ATP M.N.ADP.P N. ADP.P N.ADP + P M.N.ADP M.N + ADP, where N, M, and P represent Ncd, microtubules, and inorganic phosphate respectively, with k(+1) = 2.3 microM(-1) s(-1), k(+2) =23 s(-1), k(+3) =13 s(-1), k(+5)= 0.7 microM(-)(1) s(-)(1), and k(+6) = 3.7 s(-)(1). Phosphate release (k(+4)) was not measured directly although it is assumed to be fast relative to ADP release because Ncd is purified with ADP tightly bound at the active site. ATP hydrolysis occurs at 23 s(-)(1) prior to Ncd dissociation at 13 s(-)(1). The pathway for ATP-promoted detachment (steps 1-3) of Ncd from the microtubule is comparable to kinesin's. However, there are two major differences between the mechanisms of Ncd and kinesin. In contrast to kinesin, mantADP release for Ncd at 3.7 s(-)(1) is the slowest step in the pathway and is believed to limit steady-state turnover. Additionally, the burst amplitude observed in the pre-steady-state acid quench experiments is stoichiometric, indicating that Ncd, in contrast to kinesin, is not processive for ATP hydrolysis.  相似文献   

4.
The chemical kinetic mechanism of kinesin (K) is considered by using a consensus scheme incorporating biochemically defined open, closed and trapped states. In the absence of microtubules, the dominant species is a trapped K*ADP state, which is defined by its ultra-slow release of ADP (off rate, k(off) approximately 0.002 s(-1)) and weak microtubule binding (dissociation constant, K(d) approximately 10-20 microM). Once bound, this trapped state equilibrates with a strongly binding open state that rapidly releases ADP (k(off) approximately 300 s(-1)). After ADP release, Mg*ATP binds (on rate, k(on) approximately 2 microM(-1)s(-1)) driving formation of a closed state that is defined by hydrolysis competence and by strong binding to microtubules. Hydrolysis (k(hyd) approximately 100-300 s(-1)) and phosphate release (k(off)>100 s(-1)) both occur in this microtubule-bound closed state. Phosphate release acts as a gate that controls reversion to the trapped K*ADP state, which detaches from the microtubule, completing the cycle.  相似文献   

5.
MutS protein initiates mismatch repair with recognition of a non-Watson-Crick base-pair or base insertion/deletion site in DNA, and its interactions with DNA are modulated by ATPase activity. Here, we present a kinetic analysis of these interactions, including the effects of ATP binding and hydrolysis, reported directly from the mismatch site by 2-aminopurine fluorescence. When free of nucleotides, the Thermus aquaticus MutS dimer binds a mismatch rapidly (k(ON)=3 x 10(6) M(-1) s(-1)) and forms a stable complex with a half-life of 10 s (k(OFF)=0.07 s(-1)). When one or both nucleotide-binding sites on the MutS*mismatch complex are occupied by ATP, the complex remains fairly stable, with a half-life of 5-7 s (k(OFF)=0.1-0.14 s(-1)), although MutS(ATP) becomes incapable of (re-)binding the mismatch. When one or both nucleotide-binding sites on the MutS dimer are occupied by ADP, the MutS*mismatch complex forms rapidly (k(ON)=7.3 x 10(6) M(-1) s(-1)) and also dissociates rapidly, with a half-life of 0.4 s (k(OFF)=1.7 s(-1)). Integration of these MutS DNA-binding kinetics with previously described ATPase kinetics reveals that: (a) in the absence of a mismatch, MutS in the ADP-bound form engages in highly dynamic interactions with DNA, perhaps probing base-pairs for errors; (b) in the presence of a mismatch, MutS stabilized in the ATP-bound form releases the mismatch slowly, perhaps allowing for onsite interactions with downstream repair proteins; (c) ATP-bound MutS then moves off the mismatch, perhaps as a mobile clamp facilitating repair reactions at distant sites on DNA, until ATP is hydrolyzed (or dissociates) and the protein turns over.  相似文献   

6.
Matsuno H  Furusawa H  Okahata Y 《Biochemistry》2005,44(7):2262-2270
Catalytic DNA cleavage reactions by an ATP-dependent deoxyribonuclease (DNase) from Micrococcus luteus were monitored directly with a DNA-immobilized 27-MHz quartz-crystal microbalance (QCM). The 27-MHz QCM is a very sensitive mass-measuring device in aqueous solution, as the frequency decreases linearly with increasing mass on the electrode at a nanogram level. Three steps in ATP-dependent DNA hydrolysis reactions, including (1) binding of DNase to the end of double-stranded DNA (dsDNA) on the QCM electrode (mass increase), (2) degradation of one strand of dsDNA in the 3' --> 5' direction depending on ATP (mass decrease), and (3) release of the enzyme from the nonhydrolyzed 5'-free-ssDNA (mass decrease), could be monitored stepwise from the time dependencies of QCM frequency changes. Kinetic parameters for each step were obtained as follows. The binding constant (K(a)) of DNase to the dsDNA was determined as (28 +/- 2) x 10(6) M(-)(1) (k(on) = (8.0 +/- 0.3) x 10(3) M (-)(1) s(-)(1) and k(off) = (0.29 +/-0.01) x 10(-)(3) s(-)(1)), and it decreased to (0.79 +/- 0.16) x 10(6) M(-)(1) (k'(on) = (2.3 +/- 0.2) x 10(3) M (-)(1) s(-)(1) and k'(off) = (2.9 +/- 0.1) x 10(-)(3) s(-)(1)) for the completely nonhydrolyzed 5'-free ssDNA. This is the reason the DNase bound to the dsDNA substrate can easily release from the nonhydrolyzed 5'-free-ssDNA after the complete hydrolysis of the 3' --> 5' direction of the complementary ssDNA. K(a) values depended on the DNA structures on the QCM, and the order of these values was as follows: the dsDNA having a 4-base-mismatched base-pair end (3) > the dsDNA having a 5' 15-base overhanging end (2) > the dsDNA having a blunt end (1) > the ssDNA having a 3'-free end (4) > the ssDNA having a 5'-free end (5). Thus, DNase hardly recognized the free 5' end of ssDNA. Michaelis-Menten parameters (K(m) for ATP and k(cat)) of the hydrolysis process also could be obtained, and the order of k(cat)/K(m) was as follows: the dsDNA having a blunt end (1) approximately the dsDNA having a 4-base-mismatched base-pair end (3) > the ssDNA having a free 3' end (4) > the ssDNA having a free 5' end (5). Thus, DNase could not recognize and not hydrolyze the free 5' end of ssDNA. The DNA hydrolysis reaction could be driven by dATP and GTP (purine base) as well as ATP, whereas the cleavage efficiency was very low driven with UTP, CTP (pyrimidine base), ADP, and AMP.  相似文献   

7.
Nucleotide binding properties of two vacant noncatalytic sites of thioredoxin-activated chloroplast coupling factor 1 (CF(1)) were studied. Kinetics of nucleotide binding to noncatalytic sites is described by the first-order equation that allows for two nucleotide binding sites that differ in kinetic features. Dependence of the nucleotide binding rate on nucleotide concentration suggests that tight nucleotide binding is preceded by rapid reversible binding of nucleotides. ADP binding is cooperative. The preincubation of CF(1) with Mg(2+) produces only slight effect on the rate of ADP binding and decreases the ATP binding rate. The ATP and ADP dissociation from noncatalytic sites is described by the first-order equation for similar sites with dissociation rate constants k(-2)(ADP)=1.5 x 10(-1) min(-1) and k(-2)(ATP) congruent with 10(-3) min(-1), respectively. As follows from the study, the noncatalytic sites of CF(1) are not homogeneous. One of them retains the major part of endogenous ADP after CF(1) precipitation with ammonium sulfate. Its other two sites can bind both ADP and ATP but have different kinetic parameters and different affinity for nucleotides.  相似文献   

8.
S Sadis  L E Hightower 《Biochemistry》1992,31(39):9406-9412
The mammalian 70-kilodalton heat shock cognate protein (Hsc70) is an abundant, cytosolic molecular chaperone whose interactions with protein substrates are regulated by ATP hydrolysis. In vitro, purified Hsc70 was found to have a slow, intrinsic ATPase activity in the absence of protein substrates. The addition of an unfolded protein such as apocytochrome c stimulated ATP hydrolysis 2-3-fold. In contrast, the native holoprotein, cytochrome c, did not stimulate the ATPase rate, in accord with recent observations that 70-kilodalton heat shock proteins interact selectively with unfolded proteins. Stimulation of ATP hydrolysis by apocytochrome c was due to an increase in the Vmax, with no effect on the Km for ATP. Following hydrolysis of [3H]ATP, a relatively stable [3H]ADP.Hsc70 complex was formed. Release of [3H]ADP from Hsc70 was most efficient in the presence of other nucleotides such as ADP or ATP, suggesting that ADP release occurs as an ADP/ATP exchange reaction. The loss of radiolabeled ADP from Hsc70 in the presence of exogenous nucleotides followed first-order kinetics. In the presence of nucleotides, apocytochrome c induced a 2-fold increase in the rate of ADP release from Hsc70. Moreover, rate constants of the nucleotide exchange reaction measured in the absence and presence of apocytochrome c (0.16 and 0.34 min-1, respectively) closely matched the kcat values derived from ATP hydrolysis measurements (0.15 and 0.38 min-1, respectively). The results suggest that ADP release in a rate-limiting step in the Hsc70 ATPase reaction and that unfolded proteins stimulate ATP hydrolysis by accelerating the rate of ADP/ATP exchange.  相似文献   

9.
The reaction of mushroom (Agaricus bisporus) tyrosinase with dioxygen in the presence of several o-diphenolic substrates has been studied by steady-state and transient-phase kinetics in order to elucidate the rate-limiting step and to provide new insights into the mechanism of oxidation of these substrates. A kinetic analysis has allowed for the first time the determination of individual rate constants for several of the partial reactions that comprise the catalytic cycle. Mushroom tyrosinase rapidly reacts with dioxygen with a second-order rate constant k(+8) = 2.3 x 10(7) M(-)(1) s(-)(1), which is similar to that reported for hemocyanins [(1.3 x 10(6))-(5.7 x 10(7)) M(-)(1) s(-)(1)]. Deoxytyrosinase binds dioxygen reversibly at the binuclear Cu(I) site with a dissociation constant K(D)(O)()2 = 46.6 microM, which is similar to the value (K(D)(O)()2 = 90 microM) reported for the binding of dioxygen to Octopus vulgaris deoxyhemocyanin [Salvato et al. (1998) Biochemistry 37, 14065-14077]. Transient and steady-state kinetics showed that o-diphenols such as 4-tert-butylcatechol react significantly faster with mettyrosinase (k(+2) = 9.02 x 10(6) M(-)(1) s(-)(1)) than with oxytyrosinase (k(+6) = 5.4 x 10(5) M(-)(1) s(-)(1)). This difference is interpreted in terms of differential steric and polar effects that modulate the access of o-diphenols to the active site for these two forms of the enzyme. The values of k(cat) for several o-diphenols are also consistent with steric and polar factors controlling the mobility, orientation, and thence the reactivity of substrates at the active site of tyrosinase.  相似文献   

10.
The multidrug efflux pump P-glycoprotein (Pgp) couples drug transport to ATP hydrolysis. Previously, using a synthetic library of tetramethylrosamine ( TMR) analogues, we observed significant variation in ATPase stimulation ( V m (D)). Concentrations required for half-maximal ATPase stimulation ( K m (D)) correlated with ATP hydrolysis transition-state stabilization and ATP occlusion (EC 50 (D)) at a single site. Herein, we characterize several TMR analogues that elicit modest turnover ( k cat 相似文献   

11.
Hsp90 is an ATP-dependent molecular chaperone whose mechanism is not yet understood in detail. Here, we present the first ATPase cycle for the mitochondrial member of the Hsp90 family called Trap1 (tumor necrosis factor receptor-associated protein 1). Using biochemical, thermodynamic, and rapid kinetic methods we dissected the kinetics of the nucleotide-regulated rearrangements between the open and the closed conformations. Surprisingly, upon ATP binding, Trap1 shifts predominantly to the closed conformation (70%), but, unlike cytosolic Hsp90 from yeast, this process is rather slow at 0.076 s(-1). Because reopening (0.034 s(-1)) is about ten times faster than hydrolysis (k(hyd) = 0.0039 s(-1)), which is the rate-limiting step, Trap1 is not able to commit ATP to hydrolysis. The proposed ATPase cycle was further scrutinized by a global fitting procedure that utilizes all relevant experimental data simultaneously. This analysis corroborates our model of a two-step binding mechanism of ATP followed by irreversible ATP hydrolysis and a one-step product (ADP) release.  相似文献   

12.
Catalytic reaction pathway for the mitogen-activated protein kinase ERK2   总被引:2,自引:0,他引:2  
Prowse CN  Hagopian JC  Cobb MH  Ahn NG  Lew J 《Biochemistry》2000,39(20):6258-6266
The structural, functional, and regulatory properties of the mitogen-activated protein kinases (MAP kinases) have long attracted considerable attention owing to the critical role that these enzymes play in signal transduction. While several MAP kinase X-ray crystal structures currently exist, there is by comparison little mechanistic information available to correlate the structural data with the known biochemical properties of these molecules. We have employed steady-state kinetic and solvent viscosometric techniques to characterize the catalytic reaction pathway of the MAP kinase ERK2 with respect to the phosphorylation of a protein substrate, myelin basic protein (MBP), and a synthetic peptide substrate, ERKtide. A minor viscosity effect on k(cat) with respect to the phosphorylation of MBP was observed (k(cat) = 10 +/- 2 s(-1), k(cat)(eta) = 0.18 +/- 0.05), indicating that substrate processing occurs via slow phosphoryl group transfer (12 +/- 4 s(-1)) followed by the faster release of products (56 +/- 4 s(-1)). At an MBP concentration extrapolated to infinity, no significant viscosity effect on k(cat)/K(m(ATP)) was observed (k(cat)/K(m(ATP)) = 0.2 +/- 0.1 microM(-1) s(-1), k(cat)/K(m(ATP))(eta) = -0.08 +/- 0.04), consistent with rapid-equilibrium binding of the nucleotide. In contrast, at saturating ATP, a full viscosity effect on k(cat)/K(m) for MBP was apparent (k(cat)/K(m(MBP)) = 2.4 +/- 1 microM(-1) s(-1), k(cat)/K(m(MBP))(eta) = 1.0 +/- 0.1), while no viscosity effect was observed on k(cat)/K(m) for the phosphorylation of ERKtide (k(cat)/K(m(ERKtide)) = (4 +/- 2) x 10(-3) microM(-1) s(-1), k(cat)/K(m(ERKtide))(eta) = -0.02 +/- 0.02). This is consistent with the diffusion-limited binding of MBP, in contrast to the rapid-equilibrium binding of ERKtide, to form the ternary Michaelis complex. Calculated values for binding constants show that the estimated value for K(d(MBP)) (/= 1.5 mM). The dramatically higher catalytic efficiency of MBP in comparison to that of ERKtide ( approximately 600-fold difference) is largely attributable to the slow dissociation rate of MBP (/=56 s(-1)), from the ERK2 active site.  相似文献   

13.
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   

14.
Transient kinetic analysis of biphasic, single turnover data for the reaction of 2,2'-azino-bis[3-ethylbenzthiazoline-6-sulfonic acid] (ABTS) with horseradish peroxidase (HRPC) compound II demonstrated preequilibrium binding of ABTS (k(+5) = 7.82 x 10(4) M(-)(1) s(-)(1)) prior to rate-limiting electron transfer (k(+6) = 42.1 s(-)(1)). These data were obtained using a stopped-flow method, which included ascorbate in the reaction medium to maintain a low steady-state concentration of ABTS (pseudo-first-order conditions) and to minimize absorbance changes in the Soret region due to the accumulation of ABTS cation radicals. A steady-state kinetic analysis of the reaction confirmed that the reduction of HRPC compound II by this substrate is rate-limiting in the complete peroxidase cycle. The reaction of HRPC with o-diphenols has been investigated using a chronometric method that also included ascorbate in the assay medium to minimize the effects of nonenzymic reactions involving phenol-derived radical products. This enabled the initial rates of o-diphenol oxidation at different hydrogen peroxide and o-diphenol concentrations to be determined from the lag period induced by the presence of ascorbate. The kinetic analysis resolved the reaction of HRPC compound II with o-diphenols into two steps, initial formation of an enzyme-substrate complex followed by electron transfer from the substrate to the heme. With o-diphenols that are rapidly oxidized, the heterolytic cleavage of the O-O bond of the heme-bound hydrogen peroxide (k(+2) = 2.17 x 10(3) s(-)(1)) is rate-limiting. The size and hydrophobicity of the o-diphenol substrates are correlated with their rate of binding to HRPC, while the electron density at the C-4 hydroxyl group predominantly influences the rate of electron transfer to the heme.  相似文献   

15.
IscU, a NifU-like Fe/S-escort protein, binds to and stimulates the ATPase activity of Hsc66, a hsp70-type molecular chaperone. We present evidence that stimulation arises from interactions of IscU with the substrate-binding site of Hsc66. IscU inhibited the ability of Hsc66 to suppress the aggregation of the denatured model substrate proteins rhodanese and citrate synthase, and calorimetric and surface plasmon resonance measurements showed that ATP destabilizes Hsc66.IscU complexes in a manner expected for hsp70-substrate complexes. Studies on the interaction of IscU with Hsc66 truncation mutants further showed that IscU does not bind the isolated ATPase domain of Hsc66 but does bind and stimulate a mutant containing the ATPase domain and substrate binding beta-sandwich subdomain. These results support a role for IscU as a substrate for Hsc66 and suggest a specialized function for Hsc66 in the assembly, stabilization, or transfer of Fe/S clusters formed on IscU.  相似文献   

16.
The role of nucleotide in controlling the pre-steady-state kinetics of peptide binding to the Escherichia coli 70-kDa molecular chaperone DnaK was investigated using stopped-flow fluorescence. The peptide used in this study, fVSV13 (representing amino acids 490-502 of the vesicular stomatitis virus glycoprotein), was dansylated specifically at its N-terminus. We found that (i) between 17 and 35 degrees C in the presence of ATP the second-order rate constant (k(on)) for fVSV13 binding to DnaK exhibited almost no dependence on temperature and did not deviate significantly from 3.8 x 10(5) M(-1) s(-1). In contrast, over the same temperature range in the presence of ADP, k(on) increased by a factor of 32 (7.3 x 10(4) to 2.3 x 10(6) M(-1) s(-1)); and (ii) ATP increased the apparent first-order rate constant for the release of fVSV13 from preformed DnaK-fVSV13 complexes by several orders of magnitude relative to ADP. The activation energy parameters for fVSV13 binding to and dissociation from DnaK are compared to the activation parameters for the binding of an unrelated peptide to DnaK and are also discussed in terms of an open-to-closed equilibrium in the polypeptide-binding domain. On the basis of this comparison, it is suggested that the activation entropy term deltaS++, which is related to the structure of the DnaK-bound peptide or the degree of solvation of the peptide, is a controlling factor in the kinetics of peptide binding to DnaK.  相似文献   

17.
The hscA and hscB genes of Escherichia coli encode novel chaperone and co-chaperone proteins, designated Hsc66 and Hsc20, respectively. We have overproduced and purified Hsc66 and Hsc20 in high yield in E. coli and describe their initial characterization including absorbance, fluorescence, and circular dichroism spectra. Immunoblot analyses of E. coli cultures using antisera to Hsc66 and Hsc20 raised in rabbits establish that Hsc66 and Hsc20 are constitutively expressed at levels corresponding to cell concentration approximately 20 microM and approximately 10 microM, respectively. The levels do not change appreciably following heat shock (44 degrees C), but a small increase in Hsc20 is observed following a shift to 10 degrees C. Purified Hsc66 exhibits a low intrinsic ATPase activity (approximately 0.6 min-1 at 37 degrees C), and Hsc20 was found to stimulate this activity up to 3.8-fold with half-maximal stimulation at a concentration approximately 5 microM. These findings suggest that Hsc66 and Hsc20 comprise a molecular chaperone system similar to the prokaryotic DnaK/DnaJ and eukaryotic hsp70/hsp40 systems. Sequence differences between Hsc66 and Hsc20 compared to other members of this chaperone family, however, suggest that the Hsc66/Hsc20 system will display different peptide binding specificity and that it is likely to be subject to different regulatory mechanisms. The high level of constitutive expression and the lack of a major response to temperature changes suggest that Hsc66 and Hsc20 play an important cellular role(s) under non-stress conditions.  相似文献   

18.
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   

19.
The Escherichia coli protein YjeQ represents a protein family whose members are broadly conserved in bacteria and have been shown to be indispensable to the growth of E. coli and Bacillus subtilis [Arigoni, F., et al. (1998) Nat. Biotechnol. 16, 851]. Proteins of the YjeQ family contain all sequence motifs typical of the vast class of P-loop-containing GTPases, but show a circular permutation, with a G4-G1-G3 pattern of motifs as opposed to the regular G1-G3-G4 pattern seen in most GTPases. All YjeQ family proteins display a unique domain architecture, which includes a predicted N-terminal OB-fold RNA-binding domain, the central permuted GTPase module, and a zinc knuckle-like C-terminal cysteine cluster. This domain architecture suggests a possible role for YjeQ as a regulator of translation. YjeQ was overexpressed, purified to homogeneity, and shown to contain 0.6 equiv of GDP. Steady state kinetic analyses indicated slow GTP hydrolysis, with a k(cat) of 9.4 h(-)(1) and a K(m) for GTP of 120 microM (k(cat)/K(m) = 21.7 M(-)(1) s(-)(1)). YjeQ also hydrolyzed other nucleoside triphosphates and deoxynucleotide triphosphates such as ATP, ITP, and CTP with specificity constants (k(cat)/K(m)) ranging from 0.2 to 1.0 M(-)(1) s(-)(1). Pre-steady state kinetic analysis of YjeQ revealed a burst of nucleotide hydrolysis for GTP described by a first-order rate constant of 100 s(-)(1) as compared to a burst rate of 0.2 s(-)(1) for ATP. In addition, a variant in the G1 motif of YjeQ (S221A) was substantially impaired for GTP hydrolysis (0.3 s(-)(1)) with a less significant impact on the steady state rate (1.8 h(-)(1)). In summary, E. coli YjeQ is an unusual, circularly permuted P-loop-containing GTPase, which catalyzes GTP hydrolysis at a rate 45 000 times greater than that of turnover.  相似文献   

20.
Bernat BA  Armstrong RN 《Biochemistry》2001,40(42):12712-12718
The fosfomycin resistance protein, FosA, catalyzes the Mn(2+)-dependent addition of glutathione to the antibiotic fosfomycin, (1R,2S)-epoxypropylphosphonic acid, rendering the antibiotic inactive. The enzyme is a homodimer of 16 kDa subunits, each of which contains a single mononuclear metal site. Stopped-flow absorbance/fluorescence spectrometry provides evidence suggesting a complex kinetic mechanism for the acquisition of Mn(2+) by apoFosA. The binding of Mn(H(2)O)(6)(2+) to apoFosA alters the UV absorption and intrinsic fluorescence characteristics of the protein sufficiently to provide sensitive spectroscopic probes of metal binding. The acquisition of metal is shown to be a multistep process involving rapid preequilibrium formation of an initial complex with release of approximately two protons (k(obsd) > or = 800 s(-1)). The initial complex either rapidly dissociates or forms an intermediate coordination complex (k > 300 s(-1)) with rapid isomerization (k > or = 20 s(-1)) to a set of tight protein-metal complexes. The observed bimolecular rate constant for formation of the intermediate coordination complex is 3 x 10(5) M(-1) s(-1). The release of Mn(2+) from the protein is slow (k approximately 10(-2) s(-1)). The kinetic results suggest a more complex chelate effect than is typically observed for metal binding to simple multidentate ligands. Although the addition of the substrate, fosfomycin, has no appreciable effect on the association kinetics of enzyme and metal, it significantly decreases the dissociation rate, suggesting that the substrate interacts directly with the metal center.  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