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 15 毫秒
1.
1. The C elevation of the compound action potential (CAP) was recorded with suction electrodes from dorsal roots of rats at 25 degrees C and toads (Bufo bufo) at 10 degrees C. The C fibre CAP had a conduction velocity of 0.5 +/- 0.07 SE M per sec (N = 10) and 0.25 +/- 0.04 M per sec (N = 8) in the rat and toad nerves respectively. 2. The depressant effect of applied drugs on the amplitude of the C fibres CAP was measured. Nerves from both species had similar sensitivities to GABA. EC50 5.0 microM +/- 0.5 SEM (N = 3) and 5.5 microM +/- 1.4 (N = 3) for the rat and toad respectively. Maximum depressant effects of GABA produced in rat and toad nerves were 35% +/- 5 SEM and 17% +/- 2.5 respectively. 3. In five out of ten of the rat nerves tested kainate had a clear depressant effect (maximum 36% +/- 4.3 SEM, EC50 6.8 microM +/- 0.9 SEM, N = 3) on the C fibre CAP. Kainate, at concentrations from 100 to 500 microM, had no effect on seven toad nerves. 4. Toad nerves were about 100 times less sensitive, than rat nerves, to capsaicin (ED50 values 430 microM +/- 190 SEM and 0.7 microM +/- 0.2 respectively, N = 4). 5. The similar sensitivity of nerves in both species to GABA and differing sensitivities to kainate and capsaicin suggests that amphibian C fibres specifically lack sensitivity to capsaicin and kainate.  相似文献   

2.
The possible role of type II (cGMP-stimulated cAMP hydrolysis) phosphodiesterase (PDE) in the accentuated antagonism of muscarinic effects on heart rate during beta-stimulation via endogenous nitric oxide (NO) was evaluated. The canine isolated sinoatrial node preparation was cross circulated with arterial blood of a support dog. The sinoatrial rate of the preparation was 96 +/- 5 beats/min (n = 16) at control. Methacholine (MCh; 0.01-1 microg) injected into the right coronary artery in a bolus fashion caused dose-dependent decreases in sinoatrial rate. Under an intra-arterial infusion of isoproterenol (1 microM), resulting in approximately 50% increase in sinoatrial rate, MCh-induced decreases were markedly augmented from -18 +/- 3% to -44 +/- 4% at 0.3 mg of MCh. When N(G)-nitro-L-arginine methyl ester (100 microM) or N(G)-monomethyl-L-arginine (100 microM) were continuously infused, the augmented MCh-induced decreases in sinoatrial rate were significantly suppressed (-29 +/- 3% or -25 +/- 3%, respectively, P < 0.01). Pretreatment with either 3-isobutyl-1-methylxanthine (IBMX; 20 microM), a non-selective PDE inhibitor, or amrinone (20 microM), a selective type III (cGMP inhibited cAMP hydrolysis) PDE inhibitor, doubled the isoproterenol-induced increase in the sinoatrial rate. However, the augmented MCh-induced decreases in sinoatrial rate were significantly depressed by IBMX (from -23 +/- 5% to -14 +/- 1%, P < 0.01) but not by amrinone (to -20 +/- 3%). These results suggest that MCh-induced accentuated antagonism in the sinoatrial node pacemaker activity can be modulated by endogenous NO via an activation of the type II cyclic GMP-stimulated cAMP PDE.  相似文献   

3.
Isothermal titration calorimetry has been used to determine thermodynamic parameters of substrate binding to the oxygenase domain of neuronal nitric oxide synthase (nNOS(oxy)) in the presence of the cofactor tetrahydrobiopterin. The intermediate N(omega)-hydroxy-L-arginine (NHA) has a larger affinity than L-Arginine (L-Arg) for nNOS(oxy), with K(d)=0.4+/-0.1 microM and 1.7+/-0.3 microM at 25 degrees C, respectively. nNOS(oxy) binds NHA and L-Arg with DeltaH -4.1+/-0.2 and -1.0+/-0.1 kcal/mol and DeltaS=15 and 23 cal/Kmol respectively. NHA binding is more exothermic probably due to formation of an extra hydrogen bond in the active site compared to L-Arg. The changes in heat capacity (DeltaC(p)) are relatively small for binding of both NHA and L-Arg (-53+/-18 and -95+/-23 cal/L mol, respectively), which indicates that hydrophobic interactions contribute little to binding.  相似文献   

4.
The aim of the study was to investigate the effect of iNOS expression on eNOS and nNOS functional activity in porcine cerebral arteries. iNOS was induced in pig basilar arteries using lipopolysaccharide (LPS). Arteries expressing iNOS generated NO and relaxed when challenged with L-arginine (30 microM), an effect that was reduced by treatment with dexamethasone (coincubated with LPS) and prevented by the iNOS inhibitor 1400 W (administered 10 min prior to precontraction). eNOS was activated by A23187 and was found to be impaired in arteries that had iNOS induced (A23187 1 microM relaxation: control 110+/-8%, LPS-treated 50+/-16% ; p<0.05, N=5-6). This was due mainly to reduced formation of NO by A23187 (NO concentration in response to A23187 1 microM: control 25+/-6 nM, LPS-treated 0.8+/-1.2 nM; p<0.001, N=5-6), in addition to a small reduction in the vasodilator response to the NO-donors NOC-22 and SIN-1. Cerebral vasodilation produced by stimulation of intramural nitrergic nerves was impaired in arteries that had iNOS induced, and this was reversed by 1400 W (control 23+/-4% relaxation, LPS-treated 11+/-1% relaxation, LPS plus 1400 W 10 microM treated 25+/-2% relaxation; p<0.01 for control versus LPS, N=6). It is concluded that the induction of iNOS in cerebral arteries reduces NO-mediated vasodilation initiated by eNOS and by nNOS, primarily by modulation of NO formation.  相似文献   

5.
Responsiveness to ouabain of the inotropic and chronotropic effects in rat atrial muscles during development (3-18 wks old) was examined. In spontaneously beating rat right atrial muscles, ouabain (3-30 microM) caused a potent positive inotropic effect in a concentration-dependent manner, but failed to have a chronotropic effect; at 30 microM, 78.6 +/- 3.4% (n = 14, p<0.01) in the contractile force and -1.1 +/- 2.3% (n = 14, p>0.05) in the sinus rate in 10-wk-old rats. The myocardium during development increased the responsiveness to ouabain (10 microM) by 27.6 +/- 2.1% (n = 14, p<0.01), 58.7 +/- 3.3% (n = 14, p<0.001), and 47.2 +/- 2.3% (n = 14, p<0.001) in 3-, 10-, and 18- wk-old rats, respectively. However, the response on the sinus rate was not modified in all of the developing stages. Higher frequencies of stimulation caused the more potent inotropic effect in left atrial muscles. In the experiments using a Ca2+-sensitive fluorescent dye (Fura-2), ouabain (10 and 30 microM) increased the cellular Ca2+ concentrations by 3.0 +/- 2.1% (n = 6, p>0.05) and 12.7 +/- 1.5% (n = 6, p<0.05) in 3-wk-old rats and by 13.0 +/- 2.7% (n = 6, p<0.05) and 42.9 +/- 3.1% (n = 6, p<0.01) in 18-wk-old rats, respectively. These results suggest that the ouabain-evoked response is enhanced during development (but tends to decrease from the maximum after maturing), presumably resulting from developmental degrees of cellular mechanisms such as Na+/K+ pump activity and Na+/Ca2+ exchange and is reflected by changes in the cellular Ca2+ concentration.  相似文献   

6.
A sensitive method for the assay of sparteine oxidase activity in vitro by microsomal fractions of human liver is described. The activity of sparteine oxidase was assessed by the formation of 2- and 5-dehydrosparteines, which were estimated by capillary gas chromatography with N2-FID detection. The limit of detection of the two metabolites, 2- and 5-dehydrosparteine, was 10 pmol (2.3 ng) per sample. Sparteine oxidase activity was linear with microsomal protein concentration ranging from 25 to 200 ug and with incubation times between 5 and 60 minutes. Omission of NADPH on incubation under an atmosphere of carbon monoxide inhibited formation of both metabolites, thus indicating that aforementioned metabolites arise in reaction catalyzed by cytochrome P-450. In three liver samples from humans classified as extensive (EM) metabolizers the formation of 2- and 5-dehydrosparteines was observed, 2-dehydrosparteine being the major metabolite. In these samples sparteine oxidase activity was characterised by Vmax = 136 +/- 53 pmol/min/mg and Km = 44 +/- 12 microM for 2-dehydrosparteine formation. For 5-dehydrosparteine formation the following values were obtained: Vmax = 57 +/- 18 pmol/min/mg and Km = 42 +/- 26 microM. In a liver sample from a poor metabolizer (PM) only the formation of 2-dehydrosparteine was detected with the method of analysis used. In this sample a great increase in Km (Km PM = 3033 microM) was noted, while Vmax was very similar to those obtained for 2-dehydrosparteine formation in EM subjects (Vmax PM = 147 pmol min/mg).  相似文献   

7.
The impact of age on functional sensitivity to A(1)-adenosine receptor activation was studied in Langendorff-perfused hearts from young (1-2 mo) and old (12-18 mo) male Wistar rats. Adenosine mediated bradycardia in young and old hearts, with sensitivity enhanced approximately 10-fold in old [negative logarithm of EC(50) (pEC(50)) = 4.56 +/- 0.11] versus young hearts (pEC(50) = 3.70 +/- 0. 09). Alternatively, the nonmetabolized A(1) agonists N(6)-cyclohexyladenosine and (R)-N(6)-phenylisopropyladenosine were equipotent in young (pEC(50) = 7.43 +/- 0.12 and 6.61 +/- 0.19, respectively) and old hearts (pEC(50) = 7.07 +/- 0.10 and 6.80 +/- 0. 11, respectively), suggesting a role for uptake and/or catabolism in age-related changes in adenosine sensitivity. In support of this suggestion, [(3)H]-adenosine uptake was approximately twofold greater in young than in old hearts (from 3-100 microM adenosine). However, although inhibition of adenosine deaminase and adenosine transport with 10 microM erythro-9-(2-hydroxy-3-nonyl)adenine hydrochloride and 10 microM S-(4-nitrobenzyl)-6-thioinosine increased adenosine sensitivity three- to fourfold, it failed to abolish the sensitivity difference in old (pEC(50) = 4.95 +/- 0.08) versus young (pEC(50) = 4.29 +/- 0.13) hearts. Data indicate that 1) age increases functional A(1) receptor sensitivity to adenosine without altering the sensitivity of the A(1) receptor itself, and 2) age impairs adenosine transport and/or catabolism, but this does not explain differing functional sensitivity to adenosine. This increased functional sensitivity to adenosine may have physiological significance in the older heart.  相似文献   

8.
In cultured endothelial cells harvested from human umbilical vein (HUVEC) or bovine aorta (BAEC) the 30 min incubation with calcium ionophore A 23187 (1 microM) or ticlopidine (100 microM) caused an increase in nitrite generation in HUVEC from basal 227 +/- 37 to 372 +/- 60 or to 325 +/- 33 pmoles per 10(6) cells, respectively, and in BAEC from basal 182 +/- 17 to 378 +/- 18 or to 423 +/- 66 pmoles per 106 cells (n = 6), respectively. Calcium ionophore A 23187 (1 microM) or ticlopidine (100 microM) next to 30 min incubation with BAEC increased release of 6-keto-PGF 1alpha from basal level of 9.4 +/- 1.8 to 96.2 +/- 5.1 or to 99.5 +/- 10.2 pmoles per 10(6) cells, respectively. The pretreatment with aspirin (300 microM) cut down this rise to 4.2 +/- 0.1 pmoles per 10(6) cells (n = 8). Basal cytoplasmic calcium levels, [Ca2+]i, in immortalised HUVEC cell line - ECV304, HUVEC and BAEC were 47.7 +/- 3.3 nM (n = 53), 68.3 +/- 5.0 nM (n = 30) and 53.1 +/- 3.0 nM (n = 15), respectively. In these cultured endothelial cells calcium ionophore A 23187 (0.1 microM) produced net maximum rise in [Ca2+]i by 157 +/-27 nM (n = 16)[ ECV304], by 107 +/- 58 nM (n=4) [HUVEC], and by 231.0 +/- 41.3 nM (n = 8) [BAEC], respectively, while ticlopidine (30 microM) produced net maximum rise in [Ca2+]i by 30.0 +/- 3.2 nM (n=9)[ECV304], 48.8 +/- 15.6 nM (n = 4)[HUVEC] and 28.4 +/- 5.4 nM (n = 8)[BAEC], respectively. Effect of ticlopidine on [Ca2+]i was not only weaker than that of calcium A 23187 but also its maximum appeared after a lag period that was 2 3 times longer than that for A23187. In ECV304 clopidogrel at concentrations of 10, 30 and 100 microM produced maximum increment of [Ca2+]i by 16.5 +/- 3.8 nM (n = 7), 47.0 +/- 6.9 nM (n = 8) and 67.2 +/- 8.3 nM (n = 8), respectively. Incubation of BAEC with A23187 (microM), ticlopidine or clopidogrel (100 microM) for 2 h did not influence viability of cultured endothelial cells. We claim that thienopyridines, independently of their delayed anti-platelet properties ex vivo do release NO and PGI2 from cultured endothelial cells in vitro. The above endothelial action of thienopyridines might be mediated by a rise in [Ca2+]i, however, this possibility has not been proved.  相似文献   

9.
The objective of this study was to examine the role of oxygen radicals, protein kinase C (PKC), and ATP-sensitive K(+) (K(ATP)) channels in mediating flumazenil-produced preconditioning. Chick cardiomyocyte death was quantified using propidium iodide, and oxygen radical generation was assessed using 2',7'-dichlorofluorescin oxidation. Preconditioning was initiated with 10 min of ischemia followed by 10 min of reoxygenation. Alternatively, flumazenil was infused for 10 min and removed 10 min before ischemia. Flumazenil (10 microM) and preconditioning increased oxygen radicals [1,693 +/- 101 (n = 3) and 1,567 +/- 98 (n = 3), respectively, vs. 345 +/- 53 (n = 3) in control] and reduced cell death similarly [22 +/- 3% (n = 5) and 18 +/- 2% (n = 6), respectively, vs. controls 49 +/- 5% (n = 8)]. Protection and increased oxygen radicals by flumazenil were abolished by pretreatment with the antioxidant thiol reductant 2-mercaptopropionyl glycine (800 microM; 52 +/- 10%, n = 6). Specific PKC inhibitors Go-6976 (0.1 microM) and chelerythrine (2 microM), given during ischemia and reoxygenation, blocked flumazenil-produced protection (47 +/- 5%, n = 6). The PKC activator phorbol 12-myristate 13-acetate (0.2 microM), given during ischemia and reoxygenation, reduced cell death similarly to that with flumazenil [17 +/- 4% (n = 6) and 22 +/- 3% (n = 5)]. Finally, 5-hydroxydecanoate (1 mM), a selective mitochondrial K(ATP) channel antagonist given during ischemia and reoxygenation, abolished the protection of flumazenil and phorbol 12-myristate 13-acetate. Thus flumazenil mimics preconditioning to reduce cell death in cardiomyocytes. Oxygen radicals activate mitochondrial K(ATP) channels via PKC during the process.  相似文献   

10.
Cardiovascular effects of an essential oil from the aerial parts of Mentha x villosa (OEMV) were tested in rats using a combined in vivo and in vitro approach. In non-anesthetized normotensive rats, OEMV (1, 5, 10, 20, 30 mg kg(-1) body wt., i.v.) induced a significant and dose-dependent hypotension (-3 +/- 1.8%; -6 +/- 0.7%; -40 +/- 6.7%; -58 +/- 3.8%; -57 +/- 2.1%, respectively) associated with decreases in heart rate (-1 +/- 0.3%; -9 +/- 0.9%; -17 +/- 3.2%; -72 +/- 3.1%; -82 +/- 1.4%, respectively). The hypotensive and bradycardic responses evoked by OEMV were attenuated and blocke by pre-treatment of the animals with atropine (2 mg kg(-1) body wt., i.v.). In isolated rat atrial preparations, OEMV (10, 100, 300, 500 microg ml(-1)) produced concentration-related negative chronotropic and inotropic effects (IC50 value = 229 +/- 17 and 120 +/- 13 microg ml(-1), respectively). In isolated rat aortic rings, increasing concentrations of OEM (10, 100, 300, 500 microg ml(-1)) were able to antagonize the effects of phenylephrine (1 microM), prostaglandin F2alpha (10 microM) and KCl (80 mM)-induced contractions (IC50 value = 255 +/- 9, 174 +/- 4 and 165 +/- 14 microg ml(-1), respectively). The vasorelaxant activity induced by OEMV was attenuated significantly by either endothelium removal (IC50 value = 304 +/- 9 microg ml(-1)), NG-nitro L-arginine methyl ester (L-NAME) 100 microM (IC50 value=359 +/- 18 microg ml(-1)), L-NAME 300 microM (IC50 value = 488 +/- 20 microg ml(-1)) or indomethacin 10 microM (IC50 value = 334 +/- 18 microg ml(-1)). However, it was not affected by atropine 1 microM (IC50 value = 247 +/- 12 microg ml(-1)). Furthermore, the hypotensive response induced by OEMV was attenuated significantly after nitric oxide (NO) synthase blockade (L-NAME, 20 mg kg(-1) body wt., i.v.), while bradycardia was not altered. The results suggest that the hypotensive effect induced by OEMV is probably due to its direct cardiodepressant action and peripheral vasodilation, which can be attributed to both endothelium-dependent (via EDRFs, at least NO and prostacyclin) and endothelium-independent mechanisms (such as Ca2+ channel blockade).  相似文献   

11.
The cyclin dependent kinase (cdk) inhibitor NU6027, 4-cyclohexylmethoxy-5-nitroso-pyrimidine-2,6-diamine (IC(50) vs cdk1/cyclinB1=2.9+/-0.1 microM and IC(50) vs cdk2/cyclinA3=2.2+/-0.6 microM), was used as the basis for the design of a series of 4-alkoxy-2,6-diamino-5-nitrosopyrimidine derivatives. The synthesis and evaluation of 21 compounds as potential inhibitors of cyclin-dependent kinases 1 and 2 is described and the structure-activity relationships relating to NU6027 have been probed. Simple alkoxy- or cycloalkoxy-groups at the O(4)-position were tolerated, with the 4-(2-methylbutoxy)-derivative (IC(50) vs cdk1/cyclinB1=12+/-2 microM and cdk2/cyclinA3=13+/-4 microM) retaining significant activity. Substitutions at the N(6) position were not tolerated. Replacement of the 5-nitroso substituent with ketone, oxime and semicarbazone groups essentially abolished activity. However, the derivative bearing an isosteric 5-formyl group, 2,6-diamino-4-cyclohexylmethoxy-pyrimidine-5-carbaldehyde, showed modest activity (IC(50) vs cdk1/cyclinB1=35+/-3 microM and cdk2/cyclinA3=43+/-3 microM). The X-ray crystal structure of the 5-formyl compound bound to cdk2 has been determined to 2.3A resolution. The intramolecular H-bond deduced from the structure with NU6027 bound to cdk2 is not evident in the structure with the corresponding formyl compound. Thus the parent compound, 4-cyclohexylmethoxy-5-nitrosopyrimidine-2,6-diamine (NU6027), remains the optimal basis for future structure-activity studies for cyclin-dependent kinase inhibitors in this series.  相似文献   

12.
The cloned rabbit intestinal Na+/glucose cotransporter was expressed in Xenopus oocytes, and transmembrane currents associated with this transporter were monitored using a two-electrode voltage clamp. Addition of D-glucose to a Na(+)-containing solution bathing these oocytes generated a current which was blocked by phlorizin. Water-injected control oocytes did not exhibit any currents under these conditions. The magnitude and shape of the currents were dependent on the extracellular glucose and Na+ concentrations and the membrane potential. At Vhold = -50 mV, the Km values for glucose and Na+ were 14 +/- 2 (N = 4) microM and 17 +/- 1 (N = 3) mM, respectively. These Km values and imax exhibited voltage dependence: increasing the membrane potential from -30 to -150 mV increased KGlcm and imax threefold and decreased KNam eightfold. The reversal potential (VR) of the phlorizin-sensitive, glucose-dependent current varied with log Nao+ (slope 46 +/- 6 [N = 9] mV). In the absence of sugar, a Na(+)-dependent, phlorizin-sensitive (Ki = 3 +/- 0.5 microM) current was detected only in RNA-injected oocytes. The amplitude of this current at -50 mV was 6 +/- 1% (N = 13) of the maximum current measured in the presence of D-glucose. The VR of this sugar-independent current varied with log Nao+ (slope 63 +/- 1 [N = 4] mV), indicating that the cotransporter may carry Na+ in the absence of sugar. We conclude that the Na+/glucose cotransporter is electrogenic and that investigations of currents associated with its operation can yield valuable insights into the mechanisms of solute translocation.  相似文献   

13.
Calcium accumulation by human erythrocyte inside-out vesicles was linear for at least 30 min in the presence of ATP. In untreated inside-out vesicles, 3.76 +/- 1.44 nmol of calcium/min/unit of acetylcholinesterase were transported, compared with 10.57 +/- 2.05 (+/- S.D.; n = 11) in those treated with calmodulin. The amount of calmodulin necessary for 50% activation of Ca2+ accumulation was 60 +/- 22 ng/ml (+/- S.D.; n = 4). The Km (Ca2+) for calmodulin-stimulated accumulation was 0.8 +/- 0.05 microM (+/- S.D.; n = 5) using Ca2+ /ethylene glycol bis(beta-aminoethyl ether)N,N,N',N'-tetraacetic acid (EGTA) buffers, or 25 microM with direct addition of unbuffered calcium. In the absence of calmodulin, these values were 0.4 and 60 microM, respectively, Km (ATP) values of 90 and 60 microM in the presence and absence of calmodulin, respectively, were measured at constant magnesium concentration (3 mM). In the presence of calmodulin, a broad pH profile is exhibited from pH 6.6 to 8.2. Maximal calcium accumulation occurs at pH 7.8. In the absence of calmodulin, the pH profile exhibits a linear upward increase from pH 7.0 to 8.2. The (Ca2+-Mg2+)-ATPase activity, measured under identical conditions, was 2.40 +/- 0.72 nmol of Pi/min/unit of acetylcholinesterase in the untreated vesicles and 11.29 +/- 2.87 nmol of Pi/min/unit of acetylcholinesterase (+/- S.D.; n = 4) in calmodulin-treated vesicles. A stoichiometry of 1.6 Ca2+/ATP hydrolyzed was determined in the absence of calmodulin; in the presence of calmodulin, this ratio was decreased to 0.94 Ca2+/ATP hydrolyzed.  相似文献   

14.
Effects of pentobarbital on the calcium current of Aplysia neurons were investigated under current- and voltage-clamp conditions using the conventional two-microelectrode technique. Pentobarbital attenuated the progressive broadening of repeated action potentials of somata, suggesting a reduction in the calcium current. When calcium ion was replaced with barium ion in the perfusing solution, in which neither sodium nor potassium ions carried transmembrane currents, the barium current (IBa) which flowed through the calcium channel of the cell membrane was generated by depolarizing pulses of several hundred milliseconds applied every 1 min from a holding potential of -50 mV. The IBa was not affected by tetrodotoxin (30 microM). The current was decreased by pentobarbital (0.1-5 mM) in a dose-dependent manner. The inhibition was much greater at a lower pH of the perfusate, indicating that the uncharged form of the agent was responsible. The voltage-dependent inactivation of the IBa proceeded with two time constants [190 +/- 21 and 2020 +/- 146 msec (N = 4) at -10 mV], both of which were shortened by adding 1 mM pentobarbital [to 120 +/- 18 and 540 +/- 51 msec (N = 4), respectively]. The IBa recovered from the inactivation with two time constants [60 +/- 7 and 871 +/- 76 msec (N = 3) at -50 mV]. The anesthetic (1 mM) prolonged both of them, to 124 +/- 20 and 1480 +/- 172 msec (N = 3), respectively, resulting in a use-dependent depression of the current at 2-Hz stimulation. Pentobarbital reduced the IBa to a greater extent when the holding potential was more positive (-30 instead of -50 mV), indicating a higher affinity of the drug to the inactivated state of the channel. These findings suggest that the attenuation of the progressive broadening of successive spikes by pentobarbital is due to a decrease in the voltage- and time-dependent calcium current, ending in depression of transmitter release from the nerve terminal.  相似文献   

15.
Bifunctional thrombin inhibitors based on the sequence of hirudin45-65   总被引:1,自引:0,他引:1  
The interaction of alpha-thrombin with the hirudin (HV1) fragment N alpha-acetyl desulfo hirudin45-65 (P51) was investigated. Kinetic analysis revealed that P51 inhibits the proteolysis of a tripeptidyl substrate with Ki = 0.72 +/- 0.13 and 0.11 +/- 0.03 microM for bovine and human alpha-thrombins, respectively. The inhibition was partially competitive, affecting substrate binding to the enzyme-inhibitor complex by a factor alpha = 2 (bovine) and alpha = 4 (human) characteristic of hyperbolic inhibitors. P51 also inhibited thrombin-induced fibrin clot formation with IC50 values of 0.94 +/- 0.20 and 0.058 +/- 0.006 microM for bovine and human alpha-thrombins, respectively. The enhanced antithrombin activity for human thrombin could be attributed to species variations in the putative auxiliary "anion" exosite since N alpha-acetyl desulfo hirudin55-65 displayed the same rank order of potency shift in a clotting assay without inhibiting the amidolytic activity of either enzyme. From these observations, a potent thrombin inhibitor was designed having modified residues corresponding to the P1 and P3 recognition sites. N alpha-Acetyl[D-Phe45, Arg47] hirudin45-65 (P53) emerged as a pure competitive inhibitor with a Ki = 2.8 +/- 0.9 nM and IC50 = 4.0 +/- 0.8 nM (human alpha-thrombin) and is designated as a "bifunctional" inhibitor. Its enhanced potency could be explained by a cooperative intramolecular interaction between the COOH-terminal domain of the inhibitor and the auxiliary exosite of thrombin on the one hand, and the modified NH2-terminal residues with the catalytic site on the other.  相似文献   

16.
To assess the hypothesis that microvascular nitric oxide (NO) is critical to maintain blood flow and solute exchange, we quantified NO production in the hamster cheek pouch in vivo, correlating it with vascular dynamics. Hamsters (100-120 g) were anesthetized and prepared for measurement of microvessel diameters by intravital microscopy, of plasma flow by isotopic sodium clearance, and of NO production by chemiluminescence. Analysis of endothelial NO synthase (eNOS) location by immunocytochemistry and subcellular fractionation revealed that eNOS was present in arterioles and venules and was 67 +/- 7% membrane bound. Basal NO release was 60.1 +/- 5.1 pM/min (n = 35), and plasma flow was 2.95 +/- 0.27 microl/min (n = 29). Local NO synthase inhibition with 30 microM N(omega)-nitro-L-arginine reduced NO production to 8.6 +/- 2.6 pmol/min (-83 +/- 5%, n = 9) and plasma flow to 1.95 +/- 0.15 microl/min (-28 +/- 12%, n = 17) within 30-45 min, in parallel with constriction of arterioles (9-14%) and venules (19-25%). The effects of N(omega)-nitro-L-arginine (10-30 microM) were proportional to basal microvascular conductance (r = 0.7, P < 0.05) and fully prevented by 1 mM L-arginine. We conclude that in this tissue, NO production contributes to 35-50% of resting microvascular conductance and plasma-tissue exchange.  相似文献   

17.
Glutathione (GSH) is an important component of antioxidant defenses in airway surface liquid (ASL), a thin layer (10-30 microm) of liquid covering the epithelial cells lining the airways of the lung. Decreased levels of ASL GSH have been reported in cystic fibrosis (CF), potentially contributing to the severe oxidative stress seen in this disease. To help investigate the role of GSH in ASL, we developed a technique suitable for analysis of GSH and its oxidized form (GSSG) in microliter samples using capillary sampling followed by capillary zone electrophoresis (CZE) analysis with conductivity detection. CZE was carried out in 100 mM CHES and 40 mM lithium hydroxide with 5 mM spermine at pH 9.1 under an applied electric field of -416 V cm(-1). To prevent any autooxidation of GSH during sample manipulations, the samples were treated with N-ethylmaleimide (50 mM) to alkylate free thiol (-SH). Under these conditions, GSH and GSSG were cleanly separated without interference from common anions (e.g. Cl(-), PO(4)(3-), HCO(3)(-), etc.) and the limit of detection for ASL analysis was 11 microM for GSH and 8 microM for GSSG (S/N=3). GSH and GSSG were also measured in rat plasma. Baseline values of 897+/-210 microM (GSH) and 215+/-61 microM (GSSG) were obtained for rat ASL (n=8), whereas 12.4+/-2.7 microM (GSH) and 14.8+/-6.7 microM (GSSG) were obtained for rat plasma (n=5).  相似文献   

18.
The effect of peptides, corresponding to sequences in the skeletal muscle dihydropyridine receptor II-III loop, on Ca(2+) release from sarcoplasmic reticulum (SR) and on ryanodine receptor (RyR) calcium release channels have been compared in preparations from normal and malignant hyperthermia (MH)-susceptible pigs. Peptide A (Thr(671)-Leu(690); 36 microM) enhanced the rate of Ca(2+) release from normal SR (SR(N)) and from SR of MH-susceptible muscle (SR(MH)) by 10 +/- 3.2 nmole/mg/min and 76 +/- 9.7 nmole/mg/min, respectively. Ca (2+) release from SR(N) or SR(MH) was not increased by control peptide NB (Gly(689)-Lys(708)). AS (scrambled A sequence; 36 microM) did not alter Ca (2+) release from SR(N), but increased release from SR(MH) by 29 +/- 4.9 nmoles/mg/min. RyR channels from MH-susceptible muscle (RyR(MH)) were up to about fourfold more strongly activated by peptide A (> or =1 nM) than normal RyR channels (RyR(N)) at -40 mV. Neither NB or AS activated RyR(N). RyR(MH) showed an approximately 1.8-fold increase in mean current with 30 microM AS. Inhibition at +40 mV was stronger in RyR(MH) and seen with peptide A (> or = 0.6 microM) and AS (> or = 0.6 microM), but not NB. These results show that the Arg(615)Cys substitution in RyR(MH) has multiple effects on RyRs. We speculate that enhanced DHPR activation of RyRs may contribute to increased Ca(2+) release from SR in MH-susceptible muscle.  相似文献   

19.
Fusarium graminearum A 3/5 possesses a high affinity system (Km = 32 +/- 8 microM; mean +/- SE) for uptake of choline, which was shown to be energy-dependent and constitutive. The maximum rate of choline uptake by this system was repressed by ammonia and glucose, showing a three-fold increase in maximum activity after nitrogen (2 h) or carbon (4 h) starvation. The system was highly specific for choline with only dimethylethanolamine (Ki = 198 +/- 29 microM), betaine aldehyde (Ki = 95 +/- 14 microM) and chlorocholine (Ki = 352 +/- 40 microM) acting as competitive inhibitors. Hemicholinium-3 acted as a mixed (non-competitive) inhibitor (KIES = 1.9 +/- 0.6 microM; KIE = 3.6 +/- 1.9 microM).  相似文献   

20.
The ability of the phorbol ester tumor promoter, PDB, to activate contraction and stimulate calcium influx was investigated in rabbit thoracic aorta. PDB caused a strong, slowly-developing sustained contraction in physiological salt solution which was concentration-related (0.01 to 10.0 microM). PDB-induced contractions (0.1 microM) in calcium-free medium were attenuated but not prevented. PDB (1.0 microM) maximally stimulated Ca influx above basal control, vehicle = 39.2 +/- 2.2; PDB 1.0 microM = 70.7 +/- 6.7 mumoles Ca/kg tissue; N = 16, p less than 0.01). These data suggest that PDB activates rabbit thoracic aorta by a combination of intracellular and extracellular calcium dependent mechanisms.  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