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 39 毫秒
1.
Divalent cations are required for two roles in prothrombin-phospholipid interaction. The first role, catalysis of a prothrombin protein transition has a reaction half-life of 100 min at 0 degrees and is a prerequisite to phospholipid binding. The binding sites required for the transition have a very low cation specificity. All di- and trivalent cations tested were effective in this role with the exception of beryllium. Barium catalyzed the transition but only at high concentrations (6.6 mM was required for half-reaction). Blood-clotting Factor X, another gamma-carboxyglutamic acid-containing protein, also undergoes a cation-catalyzed protein transition which is a prerequisite to Factor X-phospholipid binding. In both proteins, the transition can be monitored by a decrease in the protein's intrinsic fluorescence. Compared to prothrombin, the Factor X transition occurs much more rapidly, has a somewhat greater specificity for cations, and requires higher concentrations of cations. This indicates that the cation binding sites provided by gamma-carboxyglutamic acid are not completely uniform in all proteins. The second role of divalent cations in prothrombin-phospholipid interaction is in the actual protein-phospholipid binding. This interaction was studied by protein fluorescence quenching resulting from excitation energy transfer to a chromophore attached to the phospholipid membrane. Only strontium and barium satisfactorily replaced calcium in this role. A number of other cations form protein-phospholipid complexes but of the wrong structure. These cations inhibit the prothrombinase complex (Factor Xa, calcium, phospholipid, Factor V). The cation specificity for Factor X-phospholipid binding is the same as for prothrombin except that higher concentrations of cations are required. Factor Xa (generated by action of Russell's viper venom on Factor X) displayed the same calcium requirements for the protein transition and phospholipid interaction as Factor X. The cation requirements of the prothrombinase complex correlate with the cation requirements of prothrombin and Factor X-phospholipid binding. Strontium is the only cation that will singly replace calcium. Barium is ineffective alone because the concentrations required to catalyze the protein transitions cause precipitation of the phospholipid. Combination of certain other cations with barium will, however, substitute for calcium. The other cations (specifically magnesium or manganous ion) catalyze the protein transitions and barium forms the correct protein-phospholipid complexes.  相似文献   

2.
The phospholipid headgroup mobility of small unilamellar vesicles composed of different mixtures of phosphatidyl-L-serine (PS) and phosphatidylcholine is characterized by the solvent relaxation behavior of the polarity sensitive dyes 6-propionyl-2-(dimethylamino)naphthalene (Prodan) and 6-palmitoyl-2-[trimethylammoniumethyl]-methylamino]naphthalene chloride (Patman). If the PS content exceeds 10%, the addition of calcium leads to a substantial deceleration of the solvent relaxation of both dyes, indicating the formation of Ca(PS)2 complexes. Addition of prothrombin and its fragment 1 leads to a further decrease of the headgroup mobility, as explained by the binding of more than two PS-molecules by a single protein molecule. Prodan monitors the outermost region of the bilayer and it clearly distinguishes between the binding of prothrombin and its fragment 1. The deeper incalated Patman does not distinguish between both proteins. The validity of the solvent relaxation technique for the investigation of the membrane binding of peripheral proteins is demonstrated by the studies of prothrombin induced changes in the steady-state fluorescence anisotropies of 1,6-diphenyl-1,3, 5-hexatriene.  相似文献   

3.
The binding of the paramagnetic metal, Mn(II), to bovine prothrombin and the thrombin-mediated cleavage products of prothrombin, i.e. fragment 1 and the prethrombin 1 has been investigated. Analysis of the Scatchard plots of the binding data reveals that prothrombin has two high affinity Mn(II) binding sites with a Kd of 1.2 +/- 1.0 X 10(-5) M and approximately two to three lower affinity Mn(II) sites with a Kd of 1.3 +/- 1.0 X 10(-4) M. Positive cooperativity in Mn(II) binding to prothrombin was observed for the strong sites. Fragment 1, the phospholipid-binding region of prothrombin, possesses two high affinity Mn(II) sites with a Kd of 2.2 +/- 1.0 X 10(-5) M and at least two lower affinity sites with a Kd of approximately 2.5 +/- 1.0 X 10(-4) M. Positive cooperativity was not observed for the binding of Mn(II) to fragment 1. Prethrombin 1 binds one Mn(II) with a Kd of 3.2 +/- 1.0 X 10(-4) M. Using the values of free Mn(II) concentration, as determined by EPR measurements and the observed enhancements of the water proton relaxation rates at various concentrations of Mn(II) and protein, the binary enhancement values (epsilon b) of the metal-protein complexes were obtained. The extrapolated values are 11 +/- 0.4 for the initial prothrombin-binding sites, and 10 +/- 0.3 for the tight binding sites of fragment 1. The unique epsilon b value obtained for prethrombin 1 was 5.3 +/- 0.7. When Mn(II) was used in a Factor Xa-metal ion-phospholipid system for activation of prothrombin, the rate of generation of thrombin was less than or equal to 5% of that obtained when Ca(II) was employed in this activation system. Addition of Mn(II) to the same activation system containing Ca(II) resulted in a marked decrease in the rate of thrombin generation, suggesting that Mn(II) probably competes for the same sites on prothrombin as Ca(II). In agreement with this is the observation that the Mn(II) sites on prothrombin could be displaced by Ca(II) at high concentrations of Ca(II).  相似文献   

4.
Activation of prothrombin by multisquamase, the prothrombin activator from the venom of Echis multisquamatus (Central Asian sand viper), is inhibited by membranes containing negatively charged anionic phospholipids. This inhibition appears to be due to the fact that the venom activator cannot activate membrane-bound prothrombin. Initial steady state rates of prothrombin activation by multisquamase in the presence of phospholipids appeared to depend on the fraction unbound prothrombin only and this phenomenon was used to quantitate binding of prothrombin to membranes of varying phospholipid composition. In this method, the initial rate of prothrombin activation by multisquamase is measured in the absence (total prothrombin) and in the presence of a procoagulant surface (rate depending only on free prothrombin) and from the difference in activation rates the amount of membrane-bound prothrombin is calculated. The validity of the method was established by determination of the binding parameters for prothrombin binding to 100 microM phospholipid vesicles composed of 20 mole% phosphatidylserine and 80 mole% phosphatidylcholine. The binding parameters obtained were Kd=0.84 microM and n=0.021 micromoles prothrombin bound per micromole phospholipid which is in agreement with literature. Due to the nature of the measurement the method is especially suitable to quantitate binding of prothrombin at concentrations as low as 5 nM prothrombin.  相似文献   

5.
Prothrombin contains two kringle domains, a structural motif common to other plasma proteins involved in hemostasis and fibrinolysis. To determine the role of the kringle domains of prothrombin, we prepared three recombinant human prothrombin forms lacking the first kringle domain (residues 63-144; PT/delta K1), the second kringle domain (residues 144-249; PT/delta K2), or both kringle domains (63-249; PT/delta K1,2). The isolated prothrombin proteins were greater than 95% pure by SDS-polyacrylamide gel electrophoresis and were well carboxylated. PT/delta K1 displayed 50% of the specific coagulant activity of plasma prothrombin, PT/delta K2 had 10% of the specific coagulant activity, and PT/delta K1,2 was inactive. Polyclonal antibodies directed against the Ca(II)-specific conformer of prothrombin bound PT/delta K1 and PT/delta K2 with the same affinity as prothrombin, indicating that the Ca(II)-induced conformational transition does not involve sites on the prothrombin kringle domains. Gel filtration studies demonstrated that radiolabeled plasma prothrombin and all of the prothrombin kringle deletion mutants bound to phospholipid vesicles in the presence of Ca(II) but not in the presence of Mg(II) or EDTA. Relative dissociation constants of 1.10 +/- 0.75 and 0.49 +/- 0.18 microM were obtained by quasielastic light scattering for the interaction of phospholipid vesicles with plasma prothrombin and PT/delta K1, respectively. These data indicate that neither the first nor the second kringle domain contain unique sites for the interaction of prothrombin with phospholipid vesicles and are not required for prothrombin-phospholipid binding.  相似文献   

6.
The dependence on the integrity of the human protein C (PC) gamma-carboxyglutamic acid (Gla) domain on its Ca(2+)-dependent binding properties to acidic phospholipid (PL) vesicles has been examined by analysis of these interactions with recombinant (r)-PC Gla domain muteins. The concentration of Ca2+ that results in 50% saturation (C50-Ca) of PL with wild-type (wt) r-PC was not altered by more than 2-fold for the following r-variants of PC, viz. [Gla6-->Asp]r-PC, [Gla14-->Asp]r-PC, [Gla19-->Asp]r-PC, [Gla25-->Asp]r-PC, [Gla29-->Asp]r-PC, and [Gln32-->Gla]r-PC. The C50-Ca was substantially higher than that of wtr-PC for [Gla7-->Asp]r-PC (8.2 mM), [Arg15-->Leu]r-PC (4.5 mM), [Gla16-->Asp]r-PC (> 10 mM), [Gla20-->Asp]r-PC (> 10 mM), and [Gla26-->Asp]r-PC (> 10 mM), indicating that the Ca(2+)-induced conformations of these latter variants interacted poorly with these acidic PL vesicles. Titration of the PL vesicles with wtr-PC at a constant concentration of 20 mM Ca2+ leads to calculation of a concentration of PC that results in 50% saturation of the PL (C50-PC) of 0.38 microM. Essentially this same value was determined for the r-mutants, [Gla7-->Asp]r-PC and [Gln32-->Gla]r-PC. An approximate 2-fold lower C50-PC was obtained for [Gla14-->Asp]r-PC (0.14 microM), [Gla25-->Asp]r-PC (0.16 microM), and [Gla29-->Asp]r-PC (0.15 microM). Somewhat higher C50-PC values were found for [Gla6-->Asp]r-PC (1.2 microM), [Arg15-->Leu]r-PC (1.2 microM), [Gla16-->Asp]r-PC (1.2 microM), [Gla19-->Asp]r-PC (1.8 microM), [Gla20-->Asp]r-PC (1.1 microM), and [Gla26-->Asp]r-PC (1.6 microM). The results of this investigation, in conjunction with other structural and functional studies with these mutants, and the x-ray crystallographic structure of the prothrombin Gla domain-Ca2+ complex, show that Gla16 and Gla26 are the most indispensable Gla residues for maintenance of the functional Ca(2+)-dependent structure of the Gla domain of PC, whereas Gla14 is the least critical Gla residue in this regard. Of the non-Gla residues investigated, Arg15 is of great importance for maintenance of a functional Ca(2+)-dependent structure of PC, and insertion of a Gla residue at position 32, a situation that exists in the cases of some other proteins of this type, does not significantly alter these characteristics of r-PC.  相似文献   

7.
The effect of external calcium concentration ([Ca2+]o) on membrane potential-dependent calcium signals in isolated tiger salamander rod and cone photoreceptor inner segments was investigated with patch-clamp and calcium imaging techniques. Mild depolarizations led to increases in intracellular Ca2+ levels ([Ca2+]i) that were smaller when [Ca2+]o was elevated to 10 mM than when it was 3 mM, even though maximum Ca2+ conductance increased 30% with the increase in [Ca2+]o. When external calcium was lowered to 1 mM [Ca2+]o, maximum Ca2+ conductance was reduced, as expected, but the mild depolarization-induced increase in [Ca2+]i was larger than in 3 mM [Ca2+]o. In contrast, when photoreceptors were strongly depolarized, the increase in [Ca2+]i was less when [Ca2+]o was reduced. An explanation for these observations comes from an assessment of Ca2+ channel gating in voltage-clamped photoreceptors under changing conditions of [Ca2+]o. Although Ca2+ conductance increased with increasing [Ca2+]o, surface charge effects dictated large shifts in the voltage dependence of Ca2+ channel gating. Relative to the control condition (3 mM [Ca2+]o), 10 mM [Ca2+]o shifted Ca2+ channel activation 8 mV positive, reducing channel open probability over a broad range of potentials. Reducing [Ca2+]o to 1 mM reduced Ca2+ conductance but shifted Ca2+ channel activation negative by 6 mV. Thus the intracellular calcium signals reflect a balance between competing changes in gating and permeation of Ca2+ channels mediated by [Ca2+]o. In mildly depolarized cells, the [Ca2+]o-induced changes in Ca2+ channel activation proved stronger than the [Ca2+]o-induced changes in conductance. In response to the larger depolarizations caused by 80 mM [K+]o, the opposite is true, with conductance changes dominating the effects on channel activation.  相似文献   

8.
OBJECTIVE: High levels of free fatty acids have been shown to impair mechanical recovery and calcium homeostasis of isolated rat hearts following hypothermic perfusion. The objective of the present study was to investigate whether inhibition of fatty acid oxidation through activation of pyruvate dehydrogenase by millimolar concentrations of pyruvate could influence functional recovery and Ca2+ homeostasis after a hypothermic insult. METHODS: Ventricular function and myocardial calcium ([Ca]total) were measured in 3 different groups of Langendorff-perfused guinea pig hearts exposed to 40 min hypothermic (15 degrees C) perfusion, followed by 30 min rewarming at 37 degrees C. The hearts were perfused with either 11.1 mM glucose (G), glucose and 1.2 mM palmitate (GP), or glucose, palmitate and 5 mM pyruvate (GPP) as energy substrates. RESULTS: All groups showed marked elevations in [Ca]total during hypothermia (from 0.6-0.7 mumol.g dry wt-1 to 9.3-12.2 mumol.g dry wt-1 at 40 min hypothermia, P < 0.05), associated with a pronounced increase in left ventricular end-diastolic pressure (LVEDP from 0-2 to 50-60 mmHg). Following rewarming, GP-perfused hearts showed significantly lower recovery of mechanical function compared to both G- and GPP-perfused hearts (% recovery of left ventricular developed pressure: 27 +/- 8 vs. 62 +/- 3 and 62 +/- 8%, respectively, P < 0.05). The reduced mechanical recovery of GP-perfused hearts was associated with elevated [Ca]total. In separate experiments we found that addition of 1.2 mM palmitate reduced glucose oxidation ([14C]glucose) from 1.77 +/- 0.28 mumol.min-1.g dry wt-1 (G-perfused hearts) to 0.15 +/- 0.04 mumol.min-1.g dry wt-1 (GP-perfused hearts, P < 0.05), implying that fatty acids had become the major substrate for oxidative phosphorylation. Fatty acid oxidation was, however, less pronounced after further addition of 5 mM pyruvate. Thus, palmitate oxidation ([3H]palmitate) was more than 40% lower in GPP-perfused than in GP-perfused hearts (0.83 +/- 0.22 vs. 1.41 +/- 0.12 mumol.min-1.g dry wt-1, P < 0.05). CONCLUSIONS: The present results demonstrate impaired ventricular function and calcium homeostasis after hypothermia in guinea pig hearts perfused with fatty acids in addition to glucose, as compared to hearts perfused with glucose alone. Furthermore, we show that these unfavourable effects of fatty acids can be overcome by an exogenous supply of pyruvate.  相似文献   

9.
1. In smooth muscles isolated from the guinea-pig trachea, the effects of dihydropyridines, nifedipine and nicardipine on contractions produced by carbachol (Cch) were studied in normal (6 mM) and excess K+ concentration (60 mM). The tonic contraction produced by 1 microM Cch was highly dependent on the external Ca2+ concentration ([Ca2+]0) and was not significantly affected by cyclopiazonic acid or thapsigargin, Ca2+ uptake inhibitor. 2. [Ca2+]0-tension curves were steeper in the presence of 1 microM Cch (the Hill coefficient: 2.5) than in the presence of 60 mM K+ (Hill coefficient: 1.6) and their ED50 of Ca2+ was 0.16 and 0.39 mM, respectively. An increase of K+ to 60 mM in the presence of 1 microM Cch shifted the curve to the left roughly in parallel (ED50: 0.12 mM, Hill coefficient: 2.3). 3. [Ca2+]0-tension curve in the presence of 1 microM Cch was shifted to the right in parallel by nifedipine (1 microM). This was markedly potentiated by 60 mM K+ (the increase in ED50 of Ca2+ being 3 times at 6 mM and 15 times at 60 mM K+). No tension was evoked by Ca2+ up to 2.5 mM in 60 mM K+ solution containing 1 microM nifedipine but no Cch. 4. In the absence of nifedipine, Cch-induced contractions were potentiated by 60 mM K+, whereas in the presence of nifedipine, Cch-induced contractions were markedly inhibited by 60 mM K+. These mechanical changes were accompanied by an increase or a decrease in intracellular Ca2+. 5. A hypothesis is presented to explain the results which suggests that the kinetics of Ca2+ influx though a single type of pathway is modulated by membrane potential and receptor activation and that the susceptibility of the pathway to dihydropyridine blockade is closely related to the Ca2+ influx kinetics with receptor activation reducing and membrane depolarization increasing the susceptibility.  相似文献   

10.
The single-channel activity of rabbit skeletal muscle ryanodine receptor (skeletal RyR) and dog cardiac RyR was studied as a function of cytosolic [Ca2+]. The studies reveal that for both skeletal and cardiac RyRs, heterogeneous populations of channels exist, rather than a uniform behavior. Skeletal muscle RyRs displayed two extremes of behavior: 1) low-activity RyRs (LA skeletal RyRs, approximately 35% of the channels) had very low open probability (Po < 0.1) at all [Ca2+] and remained closed in the presence of Mg2+ (2 mM) and ATP (1 mM); 2) high-activity RyRs (HA skeletal RyRs) had much higher activity and displayed further heterogeneity in their Po values at low [Ca2+] (< 50 nM), and in their patterns of activation by [Ca2+]. Hill coefficients for activation (nHa) varied from 0.8 to 5.2. Cardiac RyRs, in comparison, behaved more homogeneously. Most cardiac RyRs were closed at 100 nM [Ca2+] and activated in a cooperative manner (nHa ranged from 1.6 to 5.0), reaching a high Po (> 0.6) in the presence and absence of Mg2+ and ATP. Heart RyRs were much less sensitive (10x) to inhibition by [Ca2+] than skeletal RyRs. The differential heterogeneity of heart versus skeletal muscle RyRs may reflect the modulation required for calcium-induced calcium release versus depolarization-induced Ca2+ release.  相似文献   

11.
It is shown that human lung surfactant protein (SP-A) mediates selective exchange of phospholipid probes with unlabeled phospholipid in excess vesicles in the presence of calcium and NaCl. The exchange occurs without leakage of vesicle contents, or transbilayer movement (flip-flop) of the phospholipid probes, or fusion of vesicles. Individual steps preceding the exchange are dissected by a combination of protocols, and the results are operationally interpreted in terms of a model where a calcium-dependent change in SP-A triggers aggregation of vesicles followed by probe exchange between the vesicles in contact through SP-A. The contacts remain stable in the presence of calcium; i.e., the vesicles in contact do not change their partners on the time scale of several minutes. The binding of SP-A to vesicles and the aggregation of vesicles are rapid, and the aggregation is rapidly reversed by EGTA; i.e., both the forward and reverse aggregation reactions are complete in about 1 min. The exchange rate of the various probes between aggregated vesicles below 1 mM calcium in the presence of NaCl shows selectivity, i.e., a modest dependence on the net anionic charge on vesicles and for the headgroup of the probe. Exchange with lower selectivity is seen at >2 mM Ca in the absence of NaCl. SP-A binding to vesicles does not show an absolute specificity for the phospholipid structure, but the time course of the subsequent changes does. The results suggest that SP-A contacts between phospholipid interfaces could mediate the exchange of phospholipid species (trafficking and sorting) between lung surfactant pools in the hypophase and all accessible phospholipid interfaces of the alveolar space.  相似文献   

12.
The effect of Ca2+ channel-acting drugs on bovine adrenal mitochondria Ca2+ movements was investigated. Mitochondrial Ca2+ uptake is performed by an energy-driven Ca2+ uniporter with a Km of 20.9 +/- 3.2 microM and Vmax of 148.1 +/- 7.2 nmol 45Ca2+ min-1 mg-1. Ca2+ release is performed through an Na+/Ca2+ antiporter with a Km for Na+ of 4.2 +/- 0.5 mM, a Vmax of 7.5 +/- 0.4 nmol 45Ca2+ min-1 mg-1, and a Hill coefficient of 1.4 +/- 0.2 Ca2+ efflux through the mitochondrial Na+/Ca2+ exchanger was inhibited by several dihydropyridines (nitrendipine, felodipine, nimodipine, (+)isradipine) and by the benzothiazepine diltiazem with similar potencies. In contrast, neither CGP 28392, Bay-K-8644, amlodipine, nor verapamil had any effect on Ca2+ efflux. Nitrendipine at 20 microM modified neither the Km nor the Hill coefficient for Na+, whereas the Vmax was reduced to 2.9 nmol 45Ca2+ min-1 mg-1, thus demonstrating noncompetitive modulation of the Na+/Ca2+ exchanger. None of the Ca2+ channel-acting drugs assayed at 100 microM affected Ca2+ influx through the uniporter. Ca2+ channel blockers inhibited the Na+/Ca2+ antiporter and displaced the specific binding of [3H]nitrendipine to intact mitochondria with Ki values similar to the IC50s obtained for the inhibition of the Ca2+ efflux. Ca2+ channel-acting drugs that did not inhibit the Na+/Ca2+ exchanger (amlodipine, CGP 28392, Bay-K-9644, and verapamil, at concentrations of 100 microM or higher) had no effect on [3H]nitrendipine binding. These results suggest that the adrenomedullary mitochondrial dihydropyridine receptor is associated with the Na+/Ca2+ exchanger.  相似文献   

13.
The kinetics of Na(+)-Ca2+ exchange current after a cytoplasmic Ca2+ concentration jump (achieved by photolysis of DM-nitrophen) was measured in excised giant membrane patches from guinea pig or rat heart. Increasing the cytoplasmic Ca2+ concentration from 0.5 microM in the presence of 100 mM extracellular Na+ elicits an inward current that rises with a time constant tau 1 < 50 microseconds and decays to a plateau with a time constant tau 2 = 0.65 +/- 0.18 ms (n = 101) at 21 degrees C. These current signals are suppressed by Ni2+ and dichlorobenzamil. No stationary current, but a transient inward current that rises with tau 1 < 50 microseconds and decays with tau 2 = 0.28 +/- 0.06 ms (n = 53, T = 21 degrees C) is observed if the Ca2+ concentration jump is performed under conditions that promote Ca(2+)-Ca2+ exchange (i.e., no extracellular Na+, 5 mM extracellular Ca2+). The transient and stationary inward current is not observed in the absence of extracellular Ca2+ and Na+. The application of alpha-chymotrypsin reveals the influence of the cytoplasmic regulatory Ca2+ binding site on Ca(2+)-Ca2+ and forward Na(+)-Ca2+ exchange and shows that this site regulates both the transient and stationary current. The temperature dependence of the stationary current exhibits an activation energy of 70 kj/mol for temperatures between 21 degrees C and 38 degrees C, and 138 kj/mol between 10 degrees C and 21 degrees C. For the decay time constant an activation energy of 70 kj/mol is observed in the Na(+)-Ca2+ and the Ca(2+)-Ca2+ exchange mode between 13 degrees C and 35 degrees C. The data indicate that partial reactions of the Na(+)-Ca2+ exchanger associated with Ca2+ binding and translocation are very fast at 35 degrees C, with relaxation time constants of about 6700 s-1 in the forward Na(+)-Ca2+ exchange and about 12,500 s-1 in the Ca(2+)-Ca2+ exchange mode and that net negative charge is moved during Ca2+ translocation. According to model calculations, the turnover number, however, has to be at least 2-4 times smaller than the decay rate of the transient current, and Na+ inward translocation appears to be slower than Ca2+ outward movement.  相似文献   

14.
Binding of quinacrine to phospholipids and porcine pancreatic phospholipase A2 (PLA2) was investigated using fluorescence resonance energy transfer, Langmuir films, assay for the enzymatic activity, and molecular modeling. No significant binding of this drug to the zwitterionic phosphatidylcholine was observed whereas a high affinity for acidic phospholipids was revealed by quenching of pyrene-labeled phospholipid analogues. Partial reversal of this binding was observed due to the addition of 4 mM CaCl2. Quinacrine efficiently and independently of the lipid surface pressure penetrated into monolayers of phosphatidylglycerol while only a weak penetration into phosphatidylcholine films was evident. Quinacrine also bound to eosin-labeled PLA2, and the addition of 4 mM CaCl2 reversed this interaction almost completely. In the presence of acidic phospholipids both the drug and the enzyme were attached to the lipid surface. Studies on the influence of quinacrine on the activity of PLA2 toward pyrene-labeled phospholipid analogues revealed that the hydrolysis of phosphatidylcholine was progressively reduced as a function of increasing [quinacrine]. At low [CaCl2] and low quinacrine:lipid molar ratios (<1:5) quinacrine enhanced slightly the rate of hydrolysis of acidic phospholipids whereas at higher drug:lipid molar ratios (>1:2) an inhibition was observed. In the presence of 1 mM CaCl2 quinacrine inhibited PLA2-catalyzed hydrolysis of phosphatidylglycerol only when the drug:lipid molar ratio exceeded 1:1. The presence of 4 mM CaCl2 abolished nearly completely the inhibition with all the substrate analogues used. Our data suggest that the inhibition of PLA2 by quinacrine is due to its binding to the enzyme. This is supported also by molecular modeling which suggested a binding site for quinacrine close to the active site and Ca2+ binding site of the enzyme. Importantly, our data indicate that quinacrine binds avidly to acidic phospholipids and their presence may influence the drug-enzyme interaction and the inhibition of the enzyme action. Accordingly, presence of quinacrine may interfere also with other processes that require the presence of acidic lipids and/or Ca2+, such as the function of the nicotinic acetylcholine receptor.  相似文献   

15.
Demineralization during a cariogenic episode is affected by storage and transport in dental plaque of ions released from enamel, and by the effect on both of plaque fluid pH and ion concentrations. To investigate this, 45Ca effusion from a condensed film of streptococci was measured at pH 7, 6 and 5, and 0-20 mmol/l calcium. Cells were loaded into effusion chambers and the appearance of 45Ca and [3H]-inulin in carrier-containing but initially tracer-free buffer was measured. Ratios of 45Ca and [3H]-inulin activity in the initial suspending solution and at equilibrium in the clearance solution, permitted calculation of extracellular volume and bound calcium. The rate of Ca appearance was proportional to the retarded diffusion coefficient (rDe), which was related to the effective diffusion coefficient (De) by: rDe = De/(1 + R) in which R is the ratio of bound to free Ca2+. The rate of Ca2+ effusion increased with calcium concentration, converging on a value of 2.8 x 10(-10) m2/sec. At low pH it reached convergence at a lower [Ca]. This demonstrates that calcium effusion is dependent on binding, so a high proportion of binding sites in plaque will reduce mineral loss in vivo. Loss of binding sites at low pH will increase mineral loss.  相似文献   

16.
We studied the effect of endogenous adenosine on the release of [3H]acetylcholine ([3H]ACh) in cultures enriched (96.4+/-0.4%) in rat cholinergic amacrine-like neurons, as determined by labeling with an antibody against choline acetyltransferase. A small population of these cells also contained GABA. Using these cultures we observed that both [3H]ACh release, which was largely Ca2+-dependent, and 45Ca2+ influx, evoked by depolarization with 50 mM KCl, were increased when adenosine A1 receptor activation was prevented by removal of endogenous adenosine with adenosine deaminase, or by application of the A1 receptor antagonist DPCPX. Our results indicate that, in cultured rat amacrine-like neurons, the activation of A1 receptors decreases calcium influx and, thereby, inhibits [3H]ACh release.  相似文献   

17.
Treatment of cultured bovine adrenal chromaffin cells with dbcAMP increased [3H]STX binding with an EC50 of 126 microM and a half-effective time of 12 h; dbcAMP (1 mM x 18 h) raised the Bmax approximately 1.5-fold without altering the Kd value. Forskolin (0.1 mM) or IBMX (0.3 mM) also increased [3H]STX binding, while dbcGMP had no effect. Effects of dbcAMP and forskolin were abolished by H-89, an inhibitor of cAMP-dependent protein kinase. Cycloheximide (10 microgram/ml) and actinomycin D (10 microgram/ml), inhibitors of protein synthesis, nullified the stimulatory effect of dbcAMP, whereas tunicamycin, an inhibitor of protein glycosylation, had no effect. Treatment with dbcAMP augmented veratridine-induced 22Na influx, 45Ca influx via voltage-dependent Ca channels and catecholamine secretion, while the same treatment did not alter 45Ca influx and catecholamine secretion caused by high K (a direct activation of voltage-dependent Ca channels) [25]. Na influx via single Na channel calculated from 22Na influx and [3H]STX binding was quantitatively similar between non-treated and dbcAMP-treated cells. Brevetoxin allosterically enhanced veratridine-induced 22Na influx approximately 3-fold in dbcAMP-treated cells as in non-treated cells. These results suggest that cAMP-dependent protein kinase is involved in the modulation of Na channel expression in adrenal medulla.  相似文献   

18.
During synaptic transmission in the nervous system, synaptic vesicles fuse with the plasma membrane of presynaptic terminals, releasing neurotransmitter by exocytosis. The vesicle membrane is then retrieved by endocytosis and recycled into new transmitter-containing vesicles. Exocytosis in synaptic terminals is calcium-dependent, and we now report that endocytosis also is regulated by the intracellular calcium concentration ([Ca2+]i). Capacitance measurements in synaptic terminals of retinal bipolar neurons revealed that endocytosis was strongly inhibited by elevated [Ca2+]i in the range achieved by Ca(2+)-current activation. The rate of membrane retrieval was steeply dependent on [Ca2+]i, with a Hill coefficient of 4 and half-inhibition at approximately 500 nM. At [Ca2+]i > or = 900 nM, endocytosis was entirely absent. The action of internal calcium on endocytosis represents a novel negative-feedback mechanism controlling the rate of membrane recovery in synaptic terminals after neurotransmitter secretion. As membrane retrieval is the first step in vesicle recycling, this mechanism may contribute to activity-dependent synaptic depression.  相似文献   

19.
It has been suggested that the anticraving drug, acamprosate, acts via the glutamatergic system, but the exact mechanism of action is still unknown. The aim of this study was to characterize [3H]acamprosate binding and establish whether this showed any relation to sites on the NMDA receptor complex. We found saturable specific binding of [3H]acamprosate to rat brain membranes with a KD of 120 microM and a Bmax of 450 pmol/mg of protein. This acamprosate binding site was sensitive to inhibition by spermidine (IC50: 13.32 +/- 1.1 microM; Hill coefficient = 1.04), and arcaine and glutamate both potentiated the inhibitory effect of spermidine. Acamprosate binding to the acamprosate binding site was also sensitive to inhibition by divalent cations (Ca2+, Mg2+, and Sr2+). Conversely, acamprosate displaced [14C]spermidine binding from rat brain membranes with an IC50 of 645 microM and a Hill coefficient = 1.74. This inhibitory effect of acamprosate was not affected by arcaine, and was associated with a significant reduction in Bmax and binding affinity for spermidine, suggesting an allosteric interaction between acamprosate and a spermidine binding site. These data are consistent with an effect of acamprosate on the NMDA receptor protein complex, and acamprosate was also found to alter binding of [3H]dizocilpine to rat brain membranes. When no agonists were present in vitro (minimal NMDA receptor activation), acamprosate markedly potentiated [3H]dizocilpine binding at concentrations in the 5 to 200 microM range. However, under conditions of maximal receptor activation (100 microM glutamate, 30 microM glycine), acamprosate only inhibited [3H]dizocilpine binding (at concentrations concentrations >100 microM). When these binding studies were performed in the presence of 1 microM spermidine, the enhancing effects of acamprosate on [3H]dizocilpine binding were inhibited. The results show that acamprosate binds to a specific spermidine-sensitive site that modulates the NMDA receptor in a complex way. Together, with data from al Quatari et al. (see next paper), this work suggests that acamprosate acts as "partial co-agonist" at the NMDA receptor, so that low concentrations enhance activation when receptor activity is low, whereas higher concentrations are inhibitory to high levels of receptor activation. This may be relevant to the clinical effects of acamprosate in alcohol-dependent patients during abstinence.  相似文献   

20.
Dissociation of Ca2+ from human interstitial collagenase induced either by chelation with EGTA or by dilution resulted in loss of enzyme activity, a red shifted emission maximum from 334 to 340 nm and quenching of protein fluorescence by 10% at 340 nm. Circular dichroism indicated that secondary structure was unaffected by EGTA. Ca2+ binding to the EGTA-treated enzyme as assessed by fluorescence was cooperative (Hill coefficient, 2.9; 50% saturation at 0.4 mM Ca2+). The dependence of catalytic activity on [Ca2+] was also cooperative (Hill coefficient, 1.7-2.0; midpoint [Ca2+], 0.2 mM). The Ca2+-reconstituted protein was indistinguishable from the untreated enzyme by activity and fluorescence measurements. These results demonstrate that removal of Ca2+ from full-length collagenase generates a catalytically incompetent, partially unfolded state with native secondary structure but altered tertiary structure characterized by exposure of at least one tryptophyl residue to a more polar environment.  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