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.
The DS (diazepam-sensitive) and DR (diazepam-resistant) lines of mice, selected on the basis of their ataxic response to diazepam, also diverge in the physiologic response of their brain γ-aminobutyric acidA (GABAA) receptors to benzodiazepines, as indicated by augmentation of GABA-mediated chloride flux. Cross-sensitivity and -resistance to other sedatives known to interact with the GABAA-receptor have also been demonstrated in DS and DR mice. Based on the finding that these mice also show cross-sensitivity and -resistance to obtundation by halothane, we predicted that their GABAA-receptors would also exhibit a differential response to halothane as assayed by an in vitro36Cl influx assay using purified brain microvesicles. Consistent with this prediction, therapeutic concentrations of halothane enhanced 1 μmol/1 GABA-gated flux with significantly greater potency in DS than in DR mice (halothane EC50 336±64 μmol/1 (S.E.M.) vs. 605±110 μmol/1, respectively, P = 0.03), but there was no difference in maximal flux enhancement between the two lines (DS 4.7±0.4 nmol·mg−1·3−1, vs. DR 4.7±0.5nmol·mg−1·3s). Halothane (500 μmol/1) also shifted the entire GABA concentration-flux relationship significantly to the left, decreasing the EC50 for GABA in both the DS and DR lines. Importantly, the shift in the GABA concentration-flux response in the presence of halothane was more pronounced in the DS mice (GABA EC50 1.8±0.4 μmol/1vs.14.7±0.9 μmol/1 without halothane) than in the DR mice (GABA EC50 4.7±0.6 μmol/1vs.14.7±0.9 μmol/1 without halothane). This effect of halothane was highly significant, both when compared to control, and between the selected lines (P < 0.001). The findings that halothane enhances GABA-gated flux and enhances GABA's channel gating potency support the hypothesis that differential enhancement of agonist-stimulated chloride permeability at GABAA receptors could be a mechanism underlying the differential obtunding potency of halothane in DS and DR mice. However, at high GABA concentrations halothane decreased maximal chloride flux, more in DS than in DR mice (P < 0.001), which is not consistent with such a mechanism.  相似文献   

2.
The impact of chronic ethanol treatment, sufficient to induce tolerance and physical dependence, on GABAA receptor function was studied in acutely isolated neurons from the medial septum/nucleus diagonal band (MS/nDB) of adult rats using whole cell, patch-clamp recordings. In ethanol-naive Controls, GABA (0.3–300 μM) induced concentration-dependent increases in Cl current with a threshold of 0.3–1 μM, a mean maximal current of 7645 ± 2148 pA at 100–300 μM, an EC50 of 11.3 ± 1.3 μM and a slope of 1.53 ±0.07. GABA-activated currents in neurons from animals receiving two weeks of ethanol liquid diet treatment did not differ significantly on any of these measures. The rate of GABAA receptor desensitization (t1/2 = 6.49 ± 1.19 s) estimated as the time required for loss of 50% of peak current during sustained application of 10 μM GABA, as well as the residual steady state current remaining following complete desensitization for controls was unchanged by chronic ethanol. The impact of chronic ethanol treatment on the GABAA receptor modulation by lanthanum and zinc which act as positive and negative allosteric modulators, respectively, was also evaluated. Test pulses of 3 μM GABA in control neurons showed maximal potentiation by 141 ± 30% at ~ 1000 μM lanthanum with an EC50 of 107 ± 34 μM and a slope of ~ 1. Lanthanum potentiation remained the same following chronic ethanol treatment. Initial estimates based on fitted concentration response curves suggested that maximal inhibition of 3 μM GABA responses by zinc at the level of 70.2 ± 8.5% in control cells was significantly increased by chronic ethanol treatment to 95.3 ± 2.5%, although the IC50 of 60.2 ± 25 μM was not changed. However, this difference was not supported by direct tests of maximal 3–10 mM zinc concentrations. These results suggest that chronic ethanol treatment, sufficient to induce tolerance and physical dependence, probably does not lead to readily detectible changes in GABAA receptor function in MS/nDB neurons.  相似文献   

3.
During postnatal development of the central nervous system (CNS), the response of GABAA receptors to its agonist undergoes maturation from depolarizing to hyperpolarizing. This switch in polarity is due to the developmental decrease of the intracellular Cl concentration in neurons. Here we show that absence of NKCC1 in P9–P13 CA3 pyramidal neurons, through genetic manipulation or through bumetanide inhibition, results in a significant increase in cell excitability. Furthermore, the pro-convulsant agent 4-aminopyridine induces seizure-like events in NKCC1-null mice but not in wild-type mice. Measurements of muscimol responses in the presence and absence of NKCC1 shows that the Na–K–2Cl cotransporter only marginally affects intracellular Cl in P9–P13 CA3 principal neurons. However, large increases in intracellular Cl are observed in CA3 pyramidal neurons following increased hyperexcitability, indicating that P9–P13 CA3 pyramidal neurons lack robust mechanisms to regulate intracellular Cl during high synaptic activity. This increase in the Cl concentration is network-driven and activity-dependent, as it is blocked by the non-NMDA glutamate receptor antagonist DNQX. We also show that expression of the outward K–Cl cotransporter, KCC2, prevents the development of hyperexcitability, as a reduction of KCC2 expression by half results in increased susceptibility to seizure under control and 4-AP conditions.  相似文献   

4.
Wang W  Wang H  Gong N  Xu TL 《Brain research bulletin》2006,70(4-6):444-449
Enhancing inhibition via gamma-aminobutyric acid type A (GABAA) receptors contributes to anesthetic-induced impairment of long-term potentiation (LTP) of excitatory synaptic transmission, which may account for general anesthesia-associated memory impairment (amnesia). The neuron-specific K+–Cl cotransporter 2 (KCC2) is necessary for fast synaptic inhibition via maintaining the low intracellular chloride concentration required for the hyperpolarizing actions of GABA via GABAA receptors. To explore a possible role of KCC2-dependent inhibition in anesthetic-induced impairment of LTP, we used field excitatory postsynaptic potentials (fEPSP) recording and immunoblotting to study the effect of propofol on LTP maintenance and KCC2 expression in CA1 region of rat hippocampal slices. We found that propofol (30 μM) not only impaired LTP expression but also prevented LTP-accompanied downregulation of KCC2 without affecting the basal transmission of glutamatergic synapses. Moreover, the recurrent inhibition in hippocampal slices was enhanced by propofol. These propofol-induced effects were completely abolished by picrotoxin, a specific GABAA receptor-chloride channel blocker. Thus, enhancement of GABAergic inhibition and suppression of neuronal excitability may account for the sustained expression of KCC2 and the impairment of LTP by propofol. Together, this study supports a novel role for KCC2 in LTP expression and gives hints to a molecular mechanism, by which anesthetics might cause impairment of LTP.  相似文献   

5.
Nafamostat mesilate (FUT-175), a synthetic serine protease inhibitor, is active against a number of the serine proteases involved in coagulation. This has been proposed as the basis of its anticoagulant activity. We investigated the reaction of Nafamostat with bovine pancreatic trypsin as a model system. It was shown to act as a time-dependent competitive inhibitor, and the inhibition constants for the binding of Nafamostat to trypsin (i.e., Ki) and the overall inhibition constants (i.e., Ki*) were calculated to be 11.5 μM and 0.4±0.14 nM, respectively. The second-order rate constant for the reaction was 4.5±0.19×105 M−1s−1, and the product released following the acylation step, 6-amidino-2-naphthol, showed mixed-type inhibition. The competitive (Kic) and uncompetitive (Kiu) inhibition constants were 14.7 μM and 19.5 μM, respectively. Formation of the acyl-enzyme intermediate was dissected into at least two steps, with rates of 0.9 s−1 and 195 s−1. The deacylation step was relatively much slower (3.2±0.19×10−5 s−1), enabling the mass spectroscopic analysis of the acyl-enzyme intermediate, which confirmed the covalent attachment of 4-guanidinobenzoic acid to trypsin. The product of the deacylation step, 4-guanidinobenzoic acid, showed no inhibition up to a concentration of 200 μM. These data strongly suggest that while Nafamostat is a potent inhibitor of trypsin, it is actually an extremely poor substrate, and that apparent inhibition is due to the competitive formation of a very stable acyl-enzyme intermediate, analogous to some other active site titrants.  相似文献   

6.
Electrically evoked sodium currents were recorded under whole-cell patch clamp from undifferentiated HCN-1A cells. Peak sodium currents had a half-maximal activation, Vm0.5, of −22.6 ± 1.0 mV with a voltage dependence, Km, of 7.28 ± 0.39 mV−1. Steady-state inactivation indicated the presence of two types of sodium channel. One type inactivated with Vh0.5 = −93.8 ± 1.2 mV and kh = −6.8 ± 0.4 mV−1. The second type of sodium channel inactivated w Vh0.5 = −44.6 ± 1.5 mV and kh = −7.3 ± 0.4 mV−1. The occurrence of each channel type varied from cell to cell and ranged from 0 to 100% of the total sodium current. No variation in the rate of inactivation was seen when the holding potential was adjusted to eliminate the more negative of the two inactivation components. Application of tetrodotoxin (TTX) or saxitoxin (STX) revealed channel types with two different affinities for each toxin. TTX blocked peak sodium conductance with apparent IC50s of 22 nM and 5.3 μM. STX was more potent, with apparent IC50s of 1.6 nM and 1.2 μM. There was no statistical correlation between toxin sensitivity and steady-state inactivation voltage, suggesting that these properties varied independently among sodium channel types.  相似文献   

7.
The electrogenic Na/K pump current (Ip) was studied in the dissociated neostriatal neurons of the rat by using the nystatin-perforated patch recording mode. The Ip was activated by external K+ in a concentration-dependent manner with an EC50 of 0.7 mM at a holding potential (VH) of −40 mV. Other monovalent cations also caused Ip and the order of potency was Tl+>K+, Rb+>NH4+, Cs+>>>Li+. The Ip decreased with membrane hyperpolarization in an external solution containing 150 mM Na+, while the Ip did not show such voltage dependency without external Na+. Ouabain showed a steady-state inhibition of Ip in a concentration- and temperature-dependent manner at a VH of −40 mV. The IC50 values at 20 and 30°C were 7.1×10−6 and 1.3×10−6 M, respectively. The decay of Ip after adding ouabain well fitted with a single exponential function. At a VH of −40 Mv, the association (k+1) and dissociation (k−1) rate constants estimated from the time constant of the current decay at 20°C were 4.0×102 s−1 M−1 and 6.3×10−3 s−1, respectively. At 30°C, k+1 increased to 2.8×103 s−1 M−1 while k−1 showed no such change with a value of 1.8×10−3 s−1. A continuous Na+ influx was demonstrated by both the Na+-dependent leakage current and tetrodotoxin-sensitive Na+ current, which resulted in the continuous activation of the Na/K pump. It was thus concluded that the Na/K pump activity was well-maintained in the dissociated rat neostriatal neurons with distinct functional properties and that the activity of the pump was tightly connected with Na+ influxes.  相似文献   

8.
The occurrence of insuling receptors and biological responses to insulin has been investigated in trypsin-dissociated fetal rat brain cells maintained in culture for 8 days. Binding of [125]insulin to brain cells in culture was time- and pH-dependent and 85–90% specific. Porcine insulin competed for [125]insulin binding in a dose-dependent manner. Unrelated polypeptides, including angiotensin II, glucagon, bovine growth hormone, and bovine prolactin did not compete for [125]insulin binding. The half-life of [125]insulin dissociation from receptors at 24°C was 15 min and a plot of ln[B/Bo] vs time suggested two dissociation rate constants of2.7 × 10−4 sec−1 and5.0 × 10−5 sec−1. Scatchard analysis of the binding data gave a curvelinear plot which may indicate negative cooperativity or the occurrence of both high affinity(Ka = 2 × 1011M−1) and low affinity(Ka = 4 × 1010M−1) sites. Of the estimated total of 4.9 × 104 binding sites per cell, 28–30% appear to be high affinity sites.

Incubation of cultures with insuling caused a time- and dose-dependent stimulation of [3H]thymidine and [3H]uridine incorporation into TCA-precipitable material. Maximum stimulation of thymidine incorporation (2–5-fold) occured 11 h after incubation with 167 nM insulin. The same concentration of insulin caused a 2.2-fold increase in [3H]uridine incorporation in 2 h. These results indicate that cells cultured from rat brain contain specific insulin receptors capable of mediating effects of insulin on macromolecular synthesis in the central nervous system.  相似文献   


9.
GABAc receptors     
γ-Aminobutyric acid (GABA) is an important neurotransmitter that mediates inhibition in the vertebrate CNS. Until recently, two receptor subtypes were known: bicuculline-sensitive GABAA and baclofen-sensitive GABAB receptors. Several lines of evidence now indicate the existence of a third class of GABA receptor, which is distinct pharmacologically from GABAA and GABAB receptors and is found predominantly in the vertebrate retina. These novel GABAC receptors are Cl pores. They are insensitive to drugs that modulate GABAA and GABAB receptors and are activated selectively bycis-4-aminocrotonic acid.  相似文献   

10.
Superfusion of slices of the dorsal zone of the lumbar enlargement of the rat spinal cord with an artificial cerebrospinal fluid allowed the collection of cholecystokinin-like material (CCKLM) whose Ca2+-dependent release could be evoked by tissue depolarization with 30 mM K+. Studies on the possible influence of GABA and related agonists on this process showed that the amino acid, the GABAA agonist, muscimol, and the GABAB agonist, baclofen, inhibited the K+-evoked release of CCKLM from the rat spinal cord in a concentration-dependent manner. Maximal inhibition did not exceed −40% with either agonist. Furthermore, the effects of GABAA and GABAB receptor stimulation were not additive. Whereas the effects of muscimol (10 μM) and baclofen (1 μM) could be completely antagonized by bicuculline (1 μM) and phaclofen (10 μM), respectively, complete blockade of the inhibition by GABA (1 μM) could only be achieved in the presence of both antagonists. These data indicate that both GABAA and GABAB receptors are involved in the negative influence of GABA onto CCK-containing neurones within the dorsal horn of the rat spinal cord. Apparently, these receptors are not located on CCK-containing neurones themselves, since the inhibitory effect of GABA on the K+-evoked release of CCKLM could be completely prevented by tetrodotoxin (1 μM). As CCK acts centrally as an endogenous opioid antagonist, such a GABA-inhibitory control of spinal CCK-containing neurones might participate in the analgesic action of the amino acid via the intrathecal route.  相似文献   

11.
Pure cultures of rat cerebral capillary endothelium have been used to study the A- and L-systems of amino acid transport. Leucine is taken up by a non-concentrative mechanism that can be saturated, and competivively inhibited by phenylalanine. Uptake is rapid, with equilibiration apparent 3–min (al experiments performed at 37 °C). The KM for transport was 83 μM ± 26(mean ±S.E.M.., n = 3 which is in good agreement with recent vi vivo report using unanesthtissed rats. Alanine was transported by a saturable, concentrative mechanism. Dependence on Na+ -ions was demonstrated by lack of specific uptake in Na+ -ffree buffer and reduced uptake after preincubation in ouabain — Na+, K+ -ATPase inhibitor. The KM was 325 μM ±88 (mean ± S.E.M., n = 3). The finding ac an active A-system transporter in vitro suggests that the cells may have lost the polarity they demonstrate in vivo. The relevance of these findings to transport of nutrients nd drugs across the blood-brain barrier is discussed.  相似文献   

12.
In the rat soleus, the frequency of miniature end-plate potentials (MEPP) did not change after application of 10−5 M of the cholinomimetic drug carbachol between 18 °C and 34 °C but decreased by 40% at physiological temperatures of 37–38 °C. The carbachol-induced decrease in MEPP frequency was not eliminated by 10−7 to 10−8 M atropine or 3 × 10−7 (+)-tubocurarine similarly as had been previously found at frog neuromuscular junction.  相似文献   

13.
Experimental support for the steady state 15O inhalation technique, as used to measure cerebral blood flow (CBF) and oxygen utilisation (CMRO2), was obtained by describing the response of the cerebral vasculature to variations in arterial PCO2 in 6 anaesthetised dogs. Measurements were made using a positron emission tomograph (ECAT II) and arterial blood sampling, during the sequential constant inhalation of C15O2 and 15O2. Values of CBF and CMRO2 were calculated for a mixture of white and grey matter, using the steady state tracer equations derived by Jones et al. (1976). The mean CMRO2 was 3.58 ± 0.81 ml O2 · 100 ml−1 · min−1, whilst the mean CBF and OER (oxygen extraction ratio) values (for an arterial PCO2 of 40 mm Hg) were 39.9 ml · 100 ml−1 · min−1 and 0.50 ± 0.06, respectively. Arterial PCO2 was varied between 20 and 150 mm Hg. CBF was found to correlate closely with arterial PCO2, resulting in a mean slope (specific reactivity) of 1.52 ± 0.38 ml · 100 ml−1 · min−1 · mm Hg−1. Pooling the flow data resulted in a linear relationship between CBF (% change) and arterial PCO2 in the range 20–70 mm Hg, with a slope (% reactivity) of 3.2% mm Hg−1 (2 P < 0.001). The oxygen extraction ratio (OER) fell with increasing values of arterial PCO2 resulting in a stable CMRO2 throughout each study. There was no correlation between CMRO2 and artificially increased CBF.

These results support and give confidence in the use of the 15O inhalation technique for measuring CBF, OER and CMRO2.  相似文献   


14.
The uptake of [3H]ACHC and [3H]GABA into cultured neurons and astrocytes was studied. [3H]ACHC uptake was less efficient than that of GABA in both cell types and Km values for ACHC uptake into neurons and astrocytes were 40.3 μM and 210.8 μM, respectively. The corresponding Vmax values were 0.321 and 0.405 nmol·min−1·mg−1 cell protein, respectively. Kinetic studies of the effects of GABA on ACHC uptake and vice versa showed that GABA is a linear competitive inhibitor of ACHC uptake in both cell types with a Ki value of 15 μM. On the other hand, ACHC turned out to be a complex inhibitor of astrocytic GABA uptake being competitive at lower concentrations and non-competitive at higher concentrations. ACHC inhibited GABA uptake into neurons competitively with a Ki of 69 μM. It is concluded that ACHC acts primarily on neuronal GABA uptake sites but its uptake is much more complicated than hitherto anticipated.  相似文献   

15.
The effects of 5-500 μM concentrations of neutral ammonium salts on the binding of ligands to components of the GABAA receptor complex were investigated. [3H]Flunitrazepam binding to the benzodiazepine receptor was enhanced by ammonium (10–500 μM), but not sodium tartrate with EC50 = 98 μM and Emax = 31%. Further increasing ammonium tartrate concentrations (500–2500 μM) decreased [3H]flunitrazepam binding to control levels. The ammonium tartrate-induced increase in [3H]flunitrazepam binding was manifested as a 50% decrease in Kd. Furthermore, GABA increased the potency of ammonium tartrate in enhancing [3H]flunitrazepam binding by 63%. [3H]Ro 15-1788 and [3H]Ro 15-4513 binding to the benzodiazepine receptor was not significantly enhanced by ammonium tartrate (Emax ≈ 13%). Ammonium tartrate also increased, then decreased the binding of 500 nM [3H]muscimol to the GABAA receptor (EC50 = 52 μM, Emax = 30%) in a concentration-dependent manner, but had no effect on [3H]SR 95-531 binding (Emax < 16%). The ammonium tartrate-induced alterations in [3H]muscimol binding were demonstrated in saturation assays as the loss of the high affinity binding site and a 27% increase in the Bmax of the low affinity binding site. These results indicate that ammonia biphasically enhances, then returns ligand binding to both the GABA and benzodiazepine receptor components of the GABAA receptor complex to control levels in a barbiturate-like fashion. This suggests that ammonia may enhance GABAergic neurotransmission at concentrations commonly encountered in hepatic failure, an event preceding the suppression of inhibitory neuronal function observed at higher ( > 1 mM) ammonia concentrations. This increase in GABAergic neurotransmission is consistent with the clinical picture of lethargy, ataxia and cognitive deficits associated with liver failure and congenital hyperammonemia.  相似文献   

16.
To determine how [Ca2+]0 affects non-synaptic epileptogenesis in the CA1 area of hippocampal slices, we compared the extracellularly recorded hyperactivity induced by ACSF containing either micromolar (‘low’-Ca2+, LC-ACSF) or nanomolar concentrations of Ca2+ (‘zero’-Ca2+, ZC-ACSF). Both solutions effectively blocked chemical synaptic transmission but spontaneous bursts developed more quickly and consistently in ZC-ACSF and were longer in duration and more frequent than those recorded in LC-ACSF. Antidromically evoked bursts were less epileptiform, i.e., they exhibited fewer population spikes (PSs), in ZC-ACSF. Increasing [Mg2+]0 or decreasing [K+]0 suppressed spontaneous LC-ACSF bursting but only decreased the intensity and frequency of bursting in ZC-ACSF. Either manipulation increased the epileptiform nature of the antidromically evoked field potential, thereby mimicking the effect of increasing [Ca2+]0 from nanomolar to micromolar levels. Bath application of 250–500 μM GABA commonly arrested spontaneous bursting in LC-ACSF. In ZC-ACSF, GABA decreased the burst frequency but paradoxically superimposed high amplitude PSs on each burst. These effects were reversed by the GABAA receptor antagonists bicuculline methiodide or picrotoxin (50–100 μM). These results indicate that simply lowering [Ca2+]0 from micromolar to nanomolar concentrations increases the burst propensity and intensity of the CA1 population and can dramatically alter responses to pharmacological agents.  相似文献   

17.
Binding of [3H]cyclohexyladenosine (CHA) to the cellular fractions and P2 subfractions of the goldfish brain was studied. The A1 receptor density was predominantly in synaptosomal membranes. In goldfish brain synaptosomes (P2), 30 mM K+ stimulated glutamate, taurine and GABA release in a Ca2+-dependent fashion, whereas the aspartate release was Ca2+-independent. Adenosine, R-phenylisopropyladenosine (R-PIA) and CHA (100 μM) inhibited K+-stimulated glutamate release (31%, 34% and 45%, respectively). All of these effects were reversed by the selective adenosine A1 receptor antagonist, 8-cyclopentyltheophylline (CPT). In the same synaptosomal preparation, K+ (30 mM) stimulated Ca2+ influx (46.8±6.8%) and this increase was completely abolished by pretreatment with 100 nM ω-conotoxin. Pretreatment with 100 μM R-PIA or 100 μM CHA, reduced the evoked increase of intra-synaptosomal Ca2+ concentration, respectively by 37.7±4.3% and 39.7±9.0%. A possible correlation between presynaptic A1 receptor inhibition of glutamate release and inhibition of calcium influx is discussed.  相似文献   

18.
Which vasoactive substances that are synthesized in vivo could induce the release of a sufficient amount of prostacyclin (PGI2) to inhibit platelet aggregation from the vascular wall was investigated in the isolated dog heart perfused by a modified method of Langendorff. Infusion of 5 μM bradykinin or 25 u/ml crude thrombin into the heart for 30 sec resulted in the transient appearance of inhibitory activity of platelet aggregation. The inhibitory activity was stable at alkaline pH but unstable at acidic pH and thermolabile. The appearance of the inhibitory activity was prevented by treatment of the coronary vessel with 30 μM indomethacin or 1 mM tranylcypromine. These results indicated that the inhibitory activity was caused by PGI2. When 25 μM acetylcholine, 25 μM noradrenaline, 25 μM isoproterenol, 10 μM adenosine triphosphate (ATP 5 μM adenosine, 1 μM angiotensin II, 25 μM histamine or 1 μM serotonin was infused for 30 sec, no inhibitory activity of platele aggregation was observed. Bradykinin (5 × 10−9 5 × 10−6 M) and purified thrombin (1 × 10−9 1 × 10−7 M) induced a dose-dependent release of PGI2 which was assayed using a radioimmunoassay for 6-keto-prostaglandin F1 (6-keto-PGF1).  相似文献   

19.
The balance between thrombin and plasmin action has been postulated to be an important determinant of thrombosis. Measurement of plasma concentrations of fibrinopeptide A (FPA), which reflect thrombin action on the NH2-terminal end of the A chain, and of Bβ 1–42 (thrombin-increasable fibrinopeptide B immunoreactivity - TIFPB) which reflect plasmin action on the NH2-terminal end of the Bβ chain have shown systematic changes in the relative concentrations of the two peptides in thrombotic states. This paper reports kinetic data for TIFPB release by plasmin using fibrinogen, fibrin I monomer, and fibrin I polymer as substrates. For fibrinogen and fibrin I monomer the data fit the Michaelis-Menten equation. Experiments were performed with human proteins in 0.15M Trisbuffered saline at pH 7.4 and at 37°C. With fibrinogen as substrate the Km was calculated to be 0.87 μM and the Vmax 3.75 × 10−5 M/min/unit of plasmin. With fibrin I monomer as the substrate the Km was calculated to be 1.25 μM and the Vmax 5.5 × 10−5 M/min/unit of plasmin. With fibrin I polymer as substrate the data did not fit the Michaelis-Menten equation but there appeared to be no dramatic differences in rates from those obtained with the other two substrates. The influence of factor XIIIa-induced cross-linking of fibrin was not examined. It is concluded from these findings that fibrinogen and non-cross-linked fibrin I are equally good substrates for plasmin cleavage of the NH2-terminal end of the Bβ chain.  相似文献   

20.
The ability of astrocytes to sequester MeHG may indicate an astrocyte-mediated role in MeHg's neurotoxicity. Hence, studies were undertaken to assess the effects of MeHg on metabolic functions in cultured astrocytes. MeHg (10−5 M) significantly inhibited the initial rate (5 min) of uptake of86RbCl, used as a tracer for K+.86RbCl uptake was also sensitive to the omission of medium Na+. MeHg (10−5 M) also markedly inhibited the initial rate of uptake (1 min) of the Na+-dependent uptake of [3H]l-glutamate. A second neurotoxin, MnCl2 (0–5 × 10−4 M), did not alter [3H]glutamate or86RbCl uptake. MeHg, but not MnCl2, also stimulated the release of intracellular86Rb+ in a dose-dependent fashion. This effect could be prevented by the administration of MeHg as the glutathione conjugate. These observations support the hypothesis that the astrocyte plasma membrane is an important target for MeHg's toxic effect and specifically that small concentrations of this organometal inhibit the ability of astrocytes to maintain a transmembrane K+ gradient. This would be expected to compromise the ability of astrocytes to control extracellular K+ either by spatial buffering or active uptake, resulting in cellular swelling. We therefore studied volume changes in astrocytes using uptake of [14C]3-O-methyl-d-glucose, in attached cells in response to exposure to MeHg. Exposure to MeHg (0–5 × 10−4 M) caused a marked increase in the cell volume that was proportional to concentrations of MeHg.  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