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 718 毫秒
1.
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   

2.
The kinetic mechanism of the nonclaret disjunctional protein (Ncd) motor was investigated using the dimer termed MC1 (residues 209-700), which has been shown to exhibit negative-end directed motility (Chandra et al., 1993). The kinetic properties are similar to those of the monomeric Ncd motor domain (Pechatnikova and Taylor, 1997). The maximum steady-state ATPase activity of 1.5 s(-1) is half as large as for the monomeric motor. Dissociation constants in the presence of nucleotides showed the same trend but with approximately a two-fold decrease in the values: K(d) values are 1.0 microM for ADP-AlF(4), 1.1 microM for ATPgammaS, 1.5 microM for ATP, 3 microM for ADP, and 10 microM for ADP-vanadate (in 25 mM NaCl, 22 degrees C). The apparent second-order rate constants for the binding of ATP and ADP to the microtubule-motor complex (MtMC1) are 2 microM(-1) s(-1). Based on measurements at high microtubule concentrations the kinetic steps were fitted to the scheme,[see text] where N refers to one head of the dimer and T, D, and P stand for ATP, ADP, and inorganic phosphate. k(1) and k(-4) are the first-order rate constants of the transition induced by the binding of mant ATP and mant ADP respectively. ADP release is the main rate-limiting step in the MtMC1 mechanism. The binding of the MC1-mant ADP complex to microtubules released less than half of the mant ADP (alternating site reactivity). The second mant ADP is only released by the binding of nucleotides that dissociate the MtMC1 complex (ATP and ADP but not AMPPNP). The apparent rate constant for dissociation of the second mant ADP is four times smaller than the first and much smaller than the rate of dissociation of MtMC1 by ATP or ADP. These results are explained by a model in which MC1.ADP is first dissociated from the microtubule by ATP, followed by rebinding to the microtubule by the ADP-containing head. Ncd may follow a different reaction pathway than does kinesin, but the differences in rate constants do not explain the opposite direction of motion. The kinetic evidence and the high ratio of motile velocity to ATPase support a nonprocessive, low duty cycle mechanism for the Ncd motor.  相似文献   

3.
Origins of reversed directionality in the ncd molecular motor.   总被引:8,自引:1,他引:7       下载免费PDF全文
The head or motor domain of the ncd (non-claret disjunctional) molecular motor is 41% identical to that of kinesin, yet moves along microtubules in the opposite direction to kinesin. We show here that despite the reversed directionality of ncd, its kinetics in solution are homologous in key respects to those of kinesin. The rate limiting step, ADP release, occurs at 0.0033 s-1 at 100 mM NaCl and is accelerated approximately 1000-fold when the motor binds to microtubules. Other reaction steps are all very fast (> 0.1 s-1) compared with ADP release, and the motor is consequently paused in the ncd.ADP state until microtubule binding occurs (Kd = 2 microM), at which point ADP release is triggered and the motor locks onto the microtubule in a rigor-like state. These data identify close functional homology between the strong binding states of kinesin and ncd, and in view of this we discuss a possible mechanism for directional reversal, in which the strong binding states of ncd and kinesin are functionally identical, but the weak binding states are biased in opposite directions.  相似文献   

4.
Mobility of taxol inside microtubules was investigated using fluorescence recovery after photobleaching on flow-aligned bundles. Bundles were made of microtubules with either GMPCPP or GTP at the exchangeable site on the tubulin dimer. Recovery times were sensitive to bundle thickness and packing, indicating that taxol molecules are able to move laterally through the bundle. The density of open binding sites along a microtubule was varied by controlling the concentration of taxol in solution for GMPCPP samples. With >63% sites occupied, recovery times were independent of taxol concentration and, therefore, inversely proportional to the microscopic dissociation rate, k(off). It was found that 10k(off)(GMPCPP) approximately equal k(off)(GTP), consistent with, but not fully accounting for, the difference in equilibrium constants for taxol on GMPCPP and GTP microtubules. With <63% sites occupied, recovery times decreased as approximately [Tax](-1/5) for both types of microtubules. We conclude that the diffusion of taxol inside the microtubule bundle is hindered by rebinding events when open sites are within approximately 7 nm of each other.  相似文献   

5.
The effects of myosin regulatory light chain (RLC) phosphorylation and strain on adenosine diphosphate (ADP) release from cross-bridges in phasic (rabbit bladder (Rbl)) and tonic (femoral artery (Rfa)) smooth muscle were determined by monitoring fluorescence transients of the novel ADP analog, 3'-deac-eda-ADP (deac-edaADP). Fluorescence transients reporting release of 3'-deac-eda-ADP were significantly faster in phasic (0.57 +/- 0.06 s(-1)) than tonic (0.29 +/- 0.03 s(-1)) smooth muscles. Thiophosphorylation of regulatory light chains increased and strain decreased the release rate approximately twofold. The calculated (k-ADP/k+ADP) dissociation constant, Kd of unstrained, unphosphorylated cross-bridges for ADP was 0.6 microM for rabbit bladder and 0.3 microM for femoral artery. The rates of ADP release from rigor bridges and reported values of Pi release (corresponding to the steady-state adenosine triphosphatase (ATPase) rate of actomyosin (AM)) from cross-bridges during a maintained isometric contraction are similar, indicating that the ADP-release step or an isomerization preceding it may be limiting the adenosine triphosphatase rate. We conclude that the strain- and dephosphorylation-dependent high affinity for and slow ADP release from smooth muscle myosin prolongs the fraction of the duty cycle occupied by strongly bound actomyosin.ADP state(s) and contributes to the high economy of force.  相似文献   

6.
Hackney DD 《Biochemistry》2002,41(13):4437-4446
Kinesin binds to microtubules with half-site ADP release to form a tethered intermediate with one attached head without nucleotide and one tethered head that retains its bound ADP. For DKH405 containing amino acid residues 1-405 of Drosophila kinesin, release of the remaining ADP from the tethered head is slow (0.05 s(-1)), but release is accelerated by added ADP or ATP. The maximum rate of ADP-stimulated dissociation of tethered DKH405 from the microtubule is approximately 12 s(-1) as determined by turbidity. Parallel measurements of ADP-stimulated release of 2'(3')-O-(N-methylanthraniloyl)-ADP (mantADP) from the tethered intermediate by fluorescence indicate that the reaction is biphasic with a fast phase that occurs at a rate that is similar to dissociation. The rate of the slow phase is dependent on the concentrations of salt and microtubules and is equal in each case to the rate for bimolecular stimulation of ADP release by microtubules as measured independently. These results are consistent with a scheme in which the fast phase, with approximately one-third of the total amplitude change, is due to ADP-stimulated release of mantADP from the tethered intermediate at approximately 6 s(-1). This direct release of mantADP continues until terminated by dissociation of DKH405 from the microtubule at approximately 12 s(-1). The majority of the amplitude change thus occurs through bimolecular recombination of DKH405.mantADP with microtubules following initial dissociation. Analysis of a simple scheme indicates that hydrolysis of ATP at the attached head before the tethered head can release its ADP and become tightly bound may be the principal limitation to processivity.  相似文献   

7.
The reactions of pyrene-labeled actin with myosin subfragment 1 (S1) and S1-ligand complexes at low ionic strength are described by the schemes [formula: see text] where M refers to a myosin head; A is actin; L is ligand; the asterisk refers to a high fluorescence state of actin; and K1 and K3 are association constants. K1 is reduced approximately 10-fold for M.ADP or M.pyrophosphate versus M alone. The rate constant of the isomerization step (k2) is 150-200 s-1 for A*M, A*M.ADP, and A*M-pyrophosphate (20 degrees C). The interaction between the ligand the actin binding sites reduces K2 from 2,000 for A*M to 50-100 for A*M.ADP and to approximately unity for A*M-pyrophosphate. The A*M.ADP state is equated with the AM'.ADP state of Sleep and Hutton (Sleep, J., A., and Hutton, R. L. (1980) Biochemistry 19, 1276-1283).  相似文献   

8.
Ca2+ and human cardiac troponin I (cTnI) peptide binding to human cardiac troponin C (cTnC) have been investigated with the use of 2D [1H,15N] HSQC NMR spectroscopy. The spectral intensity, chemical shift, and line-shape changes were analyzed to obtain the dissociation ( K(D)) and off-rate ( k(off)) constants at 30 degrees C. The results show that sites III and IV exhibit 100-fold higher Ca2+ affinity than site II ( K(D(III,IV)) approximately 0.2 microM, K(D(II)) approximately 20 microM), but site II is partially occupied before sites III and IV are saturated. The addition of the first two equivalents of Ca2+ saturates 90% of sites III and IV and 20% of site II. This suggests that the Ca2+ occupancy of all three sites may contribute to the Ca2+-dependent regulation in muscle contraction. We have determined a k(off) of 5000 s(-1) for site II Ca2+ dissociation at 30 degrees C. Such a rapid off-rate had not been previously measured. Three cTnI peptides, cTnI(34-71), cTnI(128-147), and cTnI(147-163), were titrated to Ca2+-saturated cTnC. In each case, the binding occurs with a 1:1 stoichiometry. The determined K(D) and k(off) values are 1 microM and 5 s(-1) for cTnI(34-71), 78+/-10 microM and 5000 s(-1) for cTnI(128-147), and 150+/-10 microM and 5000 s(-1) for cTnI(147-163), respectively. Thus, the dissociation of Ca2+ from site II and cTnI(128-147) and cTnI(147-163) from cTnC are rapid enough to be involved in the contraction/relaxation cycle of cardiac muscle, while that of cTnI(34-71) from cTnC may be too slow for this process.  相似文献   

9.
10.
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   

11.
Fosfomycin [(1R,2S)-epoxypropylphosphonic acid] is a simple phosphonate found to have antibacterial activity against both Gram-positive and Gram-negative microorganisms. Early resistance to the clinical use of the antibiotic was linked to a plasmid-encoded resistance protein, FosA, that catalyzes the addition of glutathione to the oxirane ring, rendering the antibiotic inactive. Subsequent studies led to the discovery of a genomically encoded homologue in the pathogen Pseudomonas aeruginosa. The proteins are Mn(II)-dependent enzymes where the metal is proposed to act as a Lewis acid stabilizing the negative charge that develops on the oxirane oxygen in the transition state. Several simple phosphonates, including the antiviral compound phosphonoformate (K(i) = 0.4 +/- 0.1 microM, K(d) approximately 0.2 microM), are shown to be inhibitors of FosA. The crystal structure of FosA from P. aeruginosa with phosphonoformate bound in the active site has been determined at 0.95 A resolution and reveals that the inhibitor forms a five-coordinate complex with the Mn(II) center with a geometry similar to that proposed for the transition state of the reaction. Binding studies show that phosphonoformate has a near-diffusion-controlled on rate (k(on) approximately 10(7)-10(8) M(-1) s(-1)) and an off rate (k(off) = 5 s(-1)) that is slower than that for fosfomycin (k(off) = 30 s(-1)). Taken together, these data suggest that the FosA-catalyzed reaction has a very early transition state and phosphonoformate acts as a minimal transition state analogue inhibitor.  相似文献   

12.
Hsc66 from Escherichia coli is a constitutively expressed hsp70 class molecular chaperone whose activity is coupled to ATP binding and hydrolysis. To better understand the mechanism and regulation of Hsc66, we investigated the kinetics of ATP hydrolysis and the interactions of Hsc66 with nucleotides. Steady-state experiments revealed that Hsc66 has a low affinity for ATP (K(m)(ATP) = 12.7 microM) compared with other hsp70 chaperones. The kinetics of nucleotide binding were determined by analyzing changes in the Hsc66 absorbance spectrum using stopped-flow methods at 23 degrees C. ATP binding results in a rapid, biphasic increase of Hsc66 absorbance at 280 nm; this is interpreted as arising from a two-step process in which ATP binding (k(a)(ATP) = 4.2 x 10(4) M(-1) s(-1), k(d)(ATP) = 1.1 s(-1)) is followed by a slow conformational change (k(conf) = 0. 1 s(-1)). Under single turnover conditions, the ATP-induced transition decays exponentially with a rate (k(decay) = 0.0013 s(-1)) similar to that observed in both steady-state and single turnover ATP hydrolysis experiments (k(hyd) = 0.0014 s(-1)). ADP binding to Hsc66 results in a monophasic transition in the absence (k(a)(ADP) = 7 x 10(5) M(-1) s(-1), k(d)(ADP) = 60 s(-1)) and presence of physiological levels of inorganic phosphate (k(a)(ADP(P(i)) = 0.28 x 10(5) M(-1) s(-1), k(d)(ADP(P(i)) = 9.1 s(-1)). These results indicate that ATP hydrolysis is the rate-limiting step under steady-state conditions and is >10(3)-fold slower than the rate of ADP/ATP exchange. Thus, in contrast to DnaK and eukaryotic forms of hsp70 that have been characterized to date, the R if T equilibrium balance for Hsc66 is shifted in favor of the low peptide affinity T state, and regulation of the reaction cycle is expected to occur at the ATP hydrolysis step rather than at nucleotide exchange.  相似文献   

13.
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   

14.
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   

15.
Kinetic mechanism of blebbistatin inhibition of nonmuscle myosin IIb   总被引:1,自引:0,他引:1  
We examined the effect of blebbistatin on the kinetic properties of nonmuscle myosin IIB subfragment 1 (NMIIB S1). Blebbistatin is a small molecule that affects cell blebbing during the process of cell division, which has been shown to decrease the myosin ATPase activity of a number of myosins [Straight et al. (2003) Science 299, 1743-1747]. The steady-state actin-activated ATPase activity of NMIIB S1 was decreased approximately 90% at 40 microM actin in the presence of blebbistatin. Stopped-flow techniques were employed to elucidate the effect of blebbistatin on the various steps of the NMIIB S1 cross-bridge cycle. Blebbistatin did not affect ATP binding and hydrolysis. Binding to actin in the presence of ADP (0.57 +/-0.08 microM(-1) s(-1)) was reduced slightly in the presence of blebbistatin (0.38 +/- 0.03 microM(-1) s(-1)), while mantADP dissociation from acto-NMIIB S1 was reduced (approximately 30%). P(i) release was blocked in the presence of blebbistatin. Accordingly, the apparent affinity of NMIIB S1 for actin in the presence of ATP was greatly reduced. Based on the above data, we surmise that blebbistatin inhibits the ATPase activity of NMIIB S1 primarily by blocking entry into the strong binding state; secondarily, it reduces the rate of ADP release. These effects are likely mediated by binding of blebbistatin within the myosin cleft that progressively closes in forming the acto-myosin rigor state.  相似文献   

16.
Myosin IIIA is specifically expressed in photoreceptors and cochlea and is important for the phototransduction and hearing processes. In addition, myosin IIIA contains a unique N-terminal kinase domain and C-terminal tail actin-binding motif. We examined the kinetic properties of baculovirus expressed human myosin IIIA containing the kinase, motor, and two IQ domains. The maximum actin-activated ATPase rate is relatively slow (k(cat) = 0.77 +/- 0.08 s(-1)), and high actin concentrations are required to fully activate the ATPase rate (K(ATPase) = 34 +/- 11 microm). However, actin co-sedimentation assays suggest that myosin III has a relatively high steady-state affinity for actin in the presence of ATP (K(actin) approximately 7 microm). The rate of ATP binding to the motor domain is quite slow both in the presence and absence of actin (K(1)k(+2) = 0.020 and 0.001 microm(-1).s(-1), respectively). The rate of actin-activated phosphate release is more than 100-fold faster (85 s(-1)) than the k(cat), whereas ADP release in the presence of actin follows a two-step mechanism (7.0 and 0.6 s(-1)). Thus, our data suggest a transition between two actomyosin-ADP states is the rate-limiting step in the actomyosin III ATPase cycle. Our data also suggest the myosin III motor spends a large fraction of its cycle in an actomyosin ADP state that has an intermediate affinity for actin (K(d) approximately 5 microm). The long lived actomyosin-ADP state may be important for the ability of myosin III to function as a cellular transporter and actin cross-linker in the actin bundles of sensory cells.  相似文献   

17.
Inhibition of dipeptidyl peptidase IV (DPP-IV) has been proposed recently as a therapeutic approach to the treatment of type 2 diabetes. N-Substituted-glycyl-2-cyanopyrrolidide compounds, typified by NVP-DPP728 (1-[[[2-[(5-cyanopyridin-2-yl)amino]ethyl]amino]acetyl]-2-cyano-(S )-p yrrolidine), inhibit degradation of glucagon-like peptide-1 (GLP-1) and thereby potentiate insulin release in response to glucose-containing meals. In the present study NVP-DPP728 was found to inhibit human DPP-IV amidolytic activity with a K(i) of 11 nM, a k(on) value of 1.3 x 10(5) M(-)(1) s(-)(1), and a k(off) of 1.3 x 10(-)(3) s(-)(1). Purified bovine kidney DPP-IV bound 1 mol/mol [(14)C]-NVP-DPP728 with high affinity (12 nM K(d)). The dissociation constant, k(off), was 1.0 x 10(-)(3) and 1.6 x 10(-)(3) s(-)(1) in the presence of 0 and 200 microM H-Gly-Pro-AMC, respectively (dissociation t(1/2) approximately 10 min). Through kinetic evaluation of DPP-IV inhibition by the D-antipode, des-cyano, and amide analogues of NVP-DPP728, it was determined that the nitrile functionality at the 2-pyrrolidine position is required, in the L-configuration, for maximal activity (K(i) of 11 nM vs K(i) values of 5.6 to >300 microM for the other analogues tested). Surprisingly, it was found that the D-antipode, despite being approximately 500-fold less potent than NVP-DPP728, displayed identical dissociation kinetics (k(off) of 1.5 x 10(-)(3) s(-)(1)). NVP-DPP728 inhibited DPP-IV in a manner consistent with a two-step inhibition mechanism. Taken together, these data suggest that NVP-DPP728 inhibits DPP-IV through formation of a novel, reversible, nitrile-dependent complex with transition state characteristics.  相似文献   

18.
Niflumic acid is widely used to inhibit Ca(2+) -activated Cl(-) channels. However, the chemical structure of niflumic acid resembles that of diphenylamine-2-carboxylate, a drug that inhibits the cystic fibrosis transmembrane conductance regulator (CFTR) Cl(-) channel. To investigate how niflumic acid inhibits CFTR Cl(-) channel, we studied recombinant wild-type human CFTR in excised inside-out membrane patches. When added to the intracellular solution, niflumic acid caused a concentration- and voltage-dependent decrease of CFTR Cl(-) current with half-maximal inhibitory concentration (K(i)) of 253 microM and Hill co-efficient of approximately 1, at -50 mV. Niflumic acid inhibition of single CFTR Cl(-) channels was characterized by a very fast, flickery block that decreased dramatically current amplitude without altering open-probability. Consistent with these data, spectral analysis of CFTR Cl(-) currents suggested that channel block by niflumic acid was described by the closed <--> open <--> blocked kinetic scheme with blocker on rate (k(on)) = 13.9 x 10(6) M(-1)s(-1), off rate (k(off))=3348 s(-1) and dissociation constant (K(d)) = 241 microM, at -50 mV. Based on these data, we tested the effects of niflumic acid on transepithelial Cl(-) secretion and cyst growth using type I MDCK epithelial cells. Niflumic acid (200 microM) inhibited cAMP-stimulated, bumetanide-sensitive short-circuit current by 55%. Moreover, the drug potently retarded cyst growth. We conclude that niflumic acid is an open-channel blocker of CFTR that inhibits Cl(-) permeation by plugging the channel pore. It or related agents might be of value in the development of new therapies for autosomal dominant polycystic kidney disease.  相似文献   

19.
The whole-cell voltage-clamp technique was used in rat cardiac myocytes to investigate the kinetics of ADP binding to phosphorylated states of Na,K-ATPase and its effects on presteady-state Na(+)-dependent charge movements by this enzyme. Ouabain-sensitive transient currents generated by Na,K-ATPase functioning in electroneutral Na(+)-Na(+) exchange mode were measured at 23 degrees C with pipette ADP concentrations ([ADP]) of up to 4.3 mM and extracellular Na(+) concentrations ([Na](o)) between 36 and 145 mM at membrane potentials (V(M)) from -160 to +80 mV. Analysis of charge-V(M) curves showed that the midpoint potential of charge distribution was shifted toward more positive V(M) both by increasing [ADP] at constant Na(+)(o) and by increasing [Na](o) at constant ADP. The total quantity of mobile charge, on the other hand, was found to be independent of changes in [ADP] or [Na](o). The presence of ADP increased the apparent rate constant for current relaxation at hyperpolarizing V(M) but decreased it at depolarizing V(M) as compared to control (no added ADP), an indication that ADP binding facilitates backward reaction steps during Na(+)-Na(+) exchange while slowing forward reactions. Data analysis using a pseudo three-state model yielded an apparent K(d) of approximately 6 mM for ADP binding to and release from the Na,K-ATPase phosphoenzyme; a value of 130 s(-1) for k(2), a rate constant that groups Na(+) deocclusion/release and the enzyme conformational transition E(1) approximately P --> E(2)-P; a value of 162 s(-1)M(-1) for k(-2), a lumped second-order V(M)-independent rate constant describing the reverse reactions; and a Hill coefficient of approximately 1 for Na(+)(o) binding to E(2)-P. The results are consistent with electroneutral release of ADP before Na(+) is deoccluded and released through an ion well. The same approach can be used to study additional charge-moving reactions and associated electrically silent steps of the Na,K-pump and other transporters.  相似文献   

20.
Cramer J  Jaeger J  Restle T 《Biochemistry》2006,45(11):3610-3619
Here we report a detailed characterization of the biochemical and kinetic properties of the hepatitis C virus (HCV, genotype-1b, J4 consensus) RNA-dependent RNA polymerase NS5B, by performing comprehensive RNA binding, nucleotide incorporation, and protein/protein oligomerization studies. By applying equilibrium fluorescence titrations, we determined a surprisingly high dissociation constant (K(d)) of approximately 250 nM for single-stranded as well as for partially double-stranded RNA. A detailed analysis of the nucleic acid binding mechanism using pre-steady-state techniques revealed the association reaction to be nearly diffusion controlled. It occurs in a single step with a second-order rate constant (k(on)) of 0.273 nM(-)(1) s(-)(1). The dissociation of the nucleic acid-polymerase complex is fast with a dissociation rate constant (k(off)) of 59.3 s(-)(1). With short, partially double-stranded RNAs, no nucleotide incorporation could be observed, while de novo RNA synthesis with short RNA templates showed nucleotide incorporation and end-to-end template switching events. Single-turnover, single-nucleotide incorporation studies (representing here the initiation and not processive polymerization) using dinucleotide primers revealed a very slow incorporation rate (k(pol)) of 0.0007 s(-)(1) and a K(d) of the binary enzyme-nucleic acid complex for the incoming ATP of 27.7 microM. Using dynamic laser light scattering, it could be shown for the first time that oligomerization of HCV NS5B is a dynamic and monovalent salt concentration dependent process. While NS5B is highly oligomeric at low salt concentrations, monomers were only observed at NaCl concentrations above 300 mM. Binding of short RNA substrates led to a further increase in oligomerization, whereas GTP did not show any effect on protein/protein interactions. Furthermore, nucleotide incorporation studies indicate the oligomerization state does not correlate with enzymatic activities as previously proposed.  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