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.
Contents of the three forms (alpha alpha, alpha gamma, and gamma gamma) of enolase isozymes and S-100 protein in superior cervical sympathetic ganglia (SCG) excised from rats were determined by the sensitive method of enzyme immunoassay, after application of various forms of stimulation, during incubation for 3 h at 37 degrees C in vitro. The amounts of the three forms of enolase isozymes and of S-100 protein in the SCG were not altered by preganglionic or postganglionic stimulation (10 Hz) or by the addition of acetylcholine (1 mM) or a high concentration of K+ (70 mM) to the incubation medium. Norepinephrine (NE; 50 microM), as well as isoproterenol (200 microM) or 3,4-dihydroxy phenylethylamine (dopamine; 200 microM), increased the ganglionic alpha alpha and alpha gamma enolase content to 1.5 to 2.0 times the control level, whereas NE tended to slightly decrease the gamma gamma enolase content. The increase in the alpha isozymes did not appear until after 2 to 3 h of incubation with this agent as a result of an increase in protein synthesis. Propranolol, an adrenergic antagonist, partly inhibited the NE-induced increase in both alpha alpha and alpha gamma enolases. NE and its agonists also considerably increased the S-100 protein level in the SCG; however, the effect developed within half an hour of incubation as a result of the conversion of the bound S-100 protein to the water-soluble form, and did not greatly increase thereafter. Cyclic AMP (1 mM) produced the same kind of increase in the ganglionic S-100 protein content as NE did.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Doxepin (DOX) and imipramine (IM) administered by close intra-arterial injection (25, 40 and 60 μg/kg) potentiated the inhibitory effect of norepinephrine (NE) on electrically-evoked postganglionic potentials in the superior cervical ganglion of the cat. Dose-response relationships indicated no significant difference between DOX and IM with regard to their effect on NE activity. Potentiation of dopamine (DA)-induced suppression of ganglionic transmission by DOX and IM (25, 40 and 60 μg/kg) was not as pronounced as the potentiation of NE activity by these two antidepressants. Significant potentiation of DA was evident only at the 40 and 60 μg/kg dose levels of DOX and IM. Dose-response relationships indicated that potentiation of DA by DOX was significantly greater than that produced by IM.  相似文献   

3.
Abstract: Noradrenergic and dopaminergic projections converge in the medial prefrontal cortex and there is evidence of an interaction between dopamine (DA) and norepinephrine (NE) terminals in this region. We have examined the influence of drugs known to alter extracellular NE on extracellular NE and DA in medial prefrontal cortex using in vivo microdialysis. Local application of the NE uptake inhibitor desipramine (1.0 µM) delivered through a microdialysis probe increased extracellular DA (+149%) as well as NE (+201%) in medial prefrontal cortex. Furthermore, desipramine potentiated the tail shock-induced increase in both extracellular DA (stress alone, +64%; stress + desipramine, +584%) and NE (stress alone, +55%; stress + desipramine, +443%). In contrast, local application of desipramine did not affect extracellular DA in striatum, indicating that this drug does not influence DA efflux directly. Local application of the α2-adrenoceptor antagonist idazoxan (0.1 or 5.0 mM) increased extracellular NE and DA in medial prefrontal cortex. Conversely, the α2-adrenoceptor agonist clonidine (0.2 mg/kg; i.p.) decreased extracellular NE and DA in medial prefrontal cortex. These results support the hypothesis that NE terminals in medial prefrontal cortex regulate extracellular DA in this region. This regulation may be achieved by mechanisms involving an action of NE on receptors that regulate DA release (heteroreceptor regulation) and/or transport of DA into noradrenergic terminals (heterotransporter regulation).  相似文献   

4.
Close intraarterial infusion of lithium chloride (2 and 4 mEq/kg) transiently suppressed evoked postganglionic potentials in the superior cervical ganglion of the cat; lower doses (0.5 and 1 mEq/kg) had no effect on transmission. Potentiation of the ganglionic inhibitory effect of norepinephrine (NE) occurred at plasma concentrations of lithium equivalent to those found to be therapeutic in man. Concurrent administration of lithium (1 mEq/kg) and doxepin (25 mcg/kg) produced greater facilitation of the ganglionic suppressant effect of NE than either lithium or doxepin alone. Rubidium chloride (0.1, 0.5 and 1 mEq/kg) produced temporary blockade of ganglionic transmission; lower doses (0.05 and 0.075 mEq/kg) did not exhibit a ganglioplegic effect. Reduction of the ganglionic inhibitory activity of NE was observed at each dose level of rubidium. Administration of doxepin (25 mcg/kg) immediately after rubidium (0.075 mEq/kg) significantly reduced the inhibitory effect of the cation on NE activity. These results suggest that, in the cat superior cervical ganglion, lithium may block NE uptake and rubidium may promote NE release.  相似文献   

5.
Abstract: These experiments investigate the effect of block, by colchicine, of fast axonal transport in the cat's cervical sympathetic trunk (CST) on the superior cervical ganglion's choline acetyltransferase (ChAT) enzyme activity, acetylcholine (ACh) content, and ACh release. Electron microscopy on the segment of the CST exposed to colchicine 1 or 4 days earlier showed disappearance of microtubules and accumulation of vesicles and smooth membrane tubules but no disruption of the axonal cytomatrix. At 4 days following colchicine treatment, the number and size of synaptic boutons per grid square in the ganglion ipsilateral to the colchicine-treated CST were similar to those in the control ganglion. At 2 and 4 days following exposure of the CST to colchicine, ChAT activity in the ipsilateral ganglion was reduced to 76 ± 8 and 54 ± 6% of control values, respectively. ACh stores in the ganglia were also reduced (to 81 ± 6% of control values at 2 days and to 51 ± 5% of control values at 4 days). Ganglionic transmission and its sensitivity to blockade by hexamethonium during 2-Hz CST stimulation were not impaired at day 4 postcolchicine. ACh release evoked by 2-Hz stimulation of colchicine-treated axons was similar to release from untreated axons, despite the decrease in the ganglionic ACh content. In contrast, ACh release evoked by 20-Hz stimulation was depressed. The amount of ACh released during 5-Hz stimulation in the presence of vesamicol by the terminals of colchicine-treated axons was similar to that released by the terminals of untreated axons. These results suggest the following conclusions: (a) Colchicine-sensitive fast axonal transport contributes significantly to maintaining ChAT stores in preganglionic axon terminals. (b) The half-life of ChAT in sympathetic preganglionic terminals is ~4 days. (c) One consequence of colchicine-induced block of axonal transport is a reduced ACh content of preganglionic nerve terminals. (d) This decrease in ACh content appears to be the result of a loss in a reserve transmitter pool, whereas the size of the readily releasable compartment is maintained.  相似文献   

6.
F. Ali  O. Tayeb  A. Attallah 《Life sciences》1985,37(19):1757-1764
To evaluate the role of catecholamines (CA) in uremia, we used “high performance” liquid chromatographic technique with electrodetection to determine plasma and brain concentration of dopamine (DA), norepinephrine (NE) and epinephrine (E) in rats with acute and chronic uremia. The results revealed a steady elevation in plasma CA (p < 0.05) in both acutely and chronically uremic rats when compared to the level of these neurotransmitters in controls. The highest changes were observed in DA and the least in NE (16.8 ± 3.2 vs. 0.5 ± 0.2 ng/ml and 93.2 ± 11.1 vs. 68.1 ± 16.3 ng/ml. There was a positive correlation between plasma CA and the duration of uremia (r = 0.97; p < 0.05). The elevations were more pronounced in acutely uremic rats than in chronically uremic rats. This was followed by a concomitant depletion in the concentration of DA, NE and E in the brain. The defects in catecholaminergic neurotransmission as evidence of dysfunction in the autonomic nervous system may contribute to the development of neuropathy.  相似文献   

7.
Active uptake of a labelled nonmetabolizable amino acid, alpha-aminoisobutyric acid (AIB), into isolated superior cervical sympathetic ganglia (SCG) excised from adult rats was considerably stimulated by the addition of either norepinephrine (NE, 50 microM) or 3,4-dihydroxyphenylethylamine (dopamine, DA, 100 microM) to the medium during aerobic incubation for 2 h at 37 degrees C. The NE-induced increase in AIB uptake was significantly antagonized by the addition of an alpha 1-adrenoceptor antagonist (prazosin, 10 microM) in SCG axotomized 1 week prior to the examination, in which most of the ganglionic neurons had degenerated and reactive proliferation of the satellite glial components was in progress. The addition of neither acetylcholine (ACh, 1 mM) plus eserine (0.1 mM) nor cyclic nucleotides (1 mM) changed the AIB uptake by the SCG. In the axotomized SCG, the NE-evoked increase in AIB uptake was much more pronounced than that of intact or denervated SCG. A kinetic study of the active AIB uptake in the SCG showed that NE produced a decrease of the Km value and an increase in the Vmax, especially in the axotomized SCG. Ganglionic Na+, K+-ATPase activity was greatly stimulated in the presence of NE, but not by ACh. These results strongly suggest that the NE-induced enhancement of active AIB uptake in the isolated SCG is occurring in glial cells rather than in neuronal cells, with a possible alteration of membrane properties for amino acid uptake and with an apparent regulation by the stimulated transport enzyme Na+, K+-ATPase.  相似文献   

8.
The present study was designed to examine the effect of sympathetic tonic activity on parasympathetic vasodilation evoked by the trigeminal-mediated reflex in the masseter muscle in urethane-anesthetized rats. Sectioning of the superior cervical sympathetic trunk (CST) ipsilaterally increased the basal level of blood flow in the masseter muscle (MBF). Electrical stimulation of the peripheral cut end of the CST for 2 min using 2-ms pulses ipsilaterally decreased in a dependent manner the intensity (0.5-10 V) and frequency (0.1-5 Hz) of the MBF. The CST stimulation for 2 min at <0.5 Hz with 5 V using 2-ms pulses seems to be comparable with the spontaneous activity in the CST fibers innervating the masseter vasculature, because this stimulation restored the basal level of the MBF to the presectioned values. Parasympathetic vasodilation evoked by electrical stimulation of the central cut end of the lingual nerve in the masseter muscle was markedly reduced by CST stimulation for 2 min with 5 V using 2-ms pulses in a frequency-dependent manner (0.5-5 Hz). Intravenous administration of phentolamine significantly reduced the vasoconstriction induced by CST stimulation in a dose-dependent manner (0.1-1 mg/kg), but pretreatment with either phentolamine or propranolol failed to affect the sympathetic inhibition of the parasympathetic vasodilation. Our results suggest that 1) excess sympathetic activity inhibits parasympathetic vasodilation in the masseter muscle, and 2) alpha- and beta-adrenoceptors do not contribute to sympathetic inhibition of parasympathetic vasodilation, and thus some other types of receptors must be involved in this response.  相似文献   

9.
An improved sensitive, simple and time-saving method for determining catecholamine (CA) in rat brain is described. The method involves isolation on boric acid gel and high-performance liquid chromatography with electrochemical detection. Boric acid gel effectively adsorbs CA at weakly alkaline pH and the over-all recoveries of 5 ng and 10 ng samples of authentic norepinephrine (NE) and dopamine (DA) added to a homogenate of rat brain were 98.9 ± 9.2% and 103.4 ± 9.3% for NE and 96.2 ± 4.6% and 99.4 ± 4.8 for DA, respectively. Intra-assay variation was 5.3% (5 ng) and 3.0% (10 ng) for NE and 4.4% (5 ng) and 3.8% (10 ng) for DA. Inter-assay variation was 7.7% (1 ng) for NE and 5.0% (1 ng) for DA. With this analytical system, the lowest amount of NE or DA detectable was 40 pg. Application of this method to determination of the DA and NE contents of rat hypothalamus during estrous cycle revealed significant increases in the turnovers of both in the proestrus stage. This method should be useful for routine determination of plasma NE and DA because it is sensitive and inexpensive.  相似文献   

10.
Abstract— The biosynthesis of dopamine (DA) from L-tyrosine (Tyr) and L-phenylalanine (Phe) was investigated using synaptosomes prepared from the striatum and olfactory tubercle of the rat. The formation of 14CO2 from either carboxyl labeled precursor occurred exclusively within the synaptosome following hydroxylation and subsequent decarboxylation. The optimum pH for the formation of DA was 6.2 and was independent of precursor and tissue source. As pH increased beyond this optimum, synthesis from Tyr declined more rapidly than that from Phe. Synthesis obeyed Michaelis–Menton kinetics when expressed as a function of the specific activity of precursor in the medium. It was characterized by an overall Km (approx 0.9 μM) which was independent of precursor and tissue source, and was considerably lower than the Kt for accumulation of precursor by synaptosomes (15.3 and 13.3 μM for Tyr and Phe, respectively). While each precursor was an uncompetitive inhibitor of DA synthesis from the opposing labeled amino acid, Tyr was a more effective inhibitor of synthesis from Phe (Ki= 1.5 μM) than was Phe an inhibitor of synthesis from Tyr (Ki= 9.2 μM). Tryptophan inhibited synthesis competitively (Ki= 15.2 and 13.2 μM for synthesis from Tyr and Phe, respectively), and DA inhibited non-competitively (Ki= 1.1 and 0.42 μM for Tyr and Phe, respectively). A model of DA synthesis within the synaptosome is presented which attempts to integrate these data. A major feature of this schema is the proposal that newly accumulated precursor does not mix rapidly with endogenous precursor pools but rather is preferentially converted to DA.  相似文献   

11.
Summary The effects of exogenous dopamine (DA), norepinephrine (NE) and epinephrine (E) on endogenous catecholamine (CA) titers and glycemia were studied with a highly specific and sensitive radioenzymatic assay (REA) in cardiac-cannulated, prespawning sea lampreys. Neither DA nor NE had a specific effect on the endogenous titers of the other two CAs, or on glycemia. In contrast, E caused a strong increase of both DA and NE at three different doses, one of which must have been in the physiological ranges. This increase may be due to direct stimulation of E on the NE and DA cells. E also caused hyperglycemia 45 min after the injection; however, this effect occurred only with unphysiologically high doses. An estimation of the disappearance rate of exogenous CAs revealed a mammalian-like speed, ranging from 3–5.5 min.Abbreviations CA catecholamines - DA dopamine - E epinephrine - NE norepinephrine - REA radioenzymatic assay  相似文献   

12.
Levels of cyclic nucleotides and ornithine decarboxylase (ODC) activity were examined following the application of various kinds of stimuli to superior cervical sympathetic ganglia (SCG), nodose ganglia, and vagus nerve fibers excised from the rat. The level of cyclic GMP in the SCG rose rapidly to about 4.5- to 7.5-fold the unstimulated control with 10 min of incubation after applications of preganglionic electrical stimulation (10 Hz), acetylcholine (ACh; 1 mM), or high extracellular K+ ( [K+]0, 70 mM). The cyclic GMP level in nodose ganglia was increased less than in the SCG by either ACh or high [K+]0 but was not affected by ACh in vagus fibers. Cyclic AMP in the SCG was also increased about 4- to 5.5-fold over the control within 10 min with the addition of ACh, norepinephrine (NE; 0.05 mM), or high [K+]0. Although NE caused a small increase in cyclic AMP, neither ACh nor high [K+]0 produced any appreciable change in nodose ganglia or vagus fibers. The ODC activity in the SCG was increased by preganglionic stimulation of 3- to 4-hr duration but not by a shorter period. A similar change in ODC activity was caused by the addition of oxotremorine (1 mM), isoproterenol (0.1 mM), NE, cyclic AMP (1 mM), or dibutyryl cyclic GMP (1 mM). The effect was exaggerated by the further addition of 3-isobutyl-1-methylxanthine (IBMX), a phosphodiesterase inhibitor. The increase in ODC activity caused by ACh was abolished by a muscarinic cholinergic antagonist, atropine (0.01 mM), and following axotomy for a week, but not by a nicotinic antagonist or by denervation in the SCG. A similar increase in ganglionic ODC activity by NE was inhibited by an adrenergic blocker, propranolol (0.01 mM), and following axtotomy for a week, but not by denervation. Cholinergic or adrenergic stimulation did not cause an increase in ODC activity in nodose ganglia or vagus fibers. These results suggest that the stimulation-induced increase in ODC activity occurs in postganglionic neurons rather than in satellite glial cells and is mediated by muscarinic cholinergic or adrenergic receptors. The process appears to involve cyclic nucleotide-mediated protein biosynthesis in the SCG.  相似文献   

13.
Pekin ducks (Anas platyrhynchos) were bilaterally adrenalectomized (biADX), injected with 1 mg of triamcinolone (TRIAM) kg bw−1 im and given 0.9% saline drinking water during a 24 h recovery period followed by chemical sympathectomy with 6OH DOPA 3 h before the start of experimental observations. Baseline plasma dopamine (DA) concentrations decreased from 283 ± 88.5 pmol ml−1 to 42.4 ± 11.1 pmol ml−1; epinephrine (E) from 142 ± 46 pmol ml−1 to 18.4 ± 9.2 pmol ml−1 and norepinephrine (NE) from 742 ± 84 pmol ml−1 to 406 ± 38 pmol ml−1 1 day after biADX + TRIAM but before chemical sympathectomy. Baseline MABP increased from 132 ± 3.2 mmHg to 209 ± 14.3 mmHg (P < 0.05) in response to TRIAM. After chemical sympathectomy with 6OH DOPA there was an additional 90% decrease in plasma NE to 42 ± 9.4 pmol ml−1 and a concurrent 60% decrease in MABP to 83.4 ± 6.9 mmHg (P < 0.05). Nasal fluid secretion was maintained by the continuous infusion of hypertonic saline (1,000 mosmol kg H2O−1 at a rate of 0.3 ml kg−1 min−1). Rates of nasal fluid secretion and fluid electrolyte concentrations were unchanged following biADX + TRIAM + 6OH DOPA. Angiotensin II (ANG II; dose 1 μg kg bw−1 i.v.), attenuated nasal fluid secretion showing that the response to ANG II was not NE- dependent. Plasma NE concentrations decreased following Tyramine i.v. (33 ± 8.5 pmol ml−1) there being no vasopressor response. This is the first report of the ANG II induced attenuation of duck salt gland secretion in the absence of measurable E and NE.  相似文献   

14.
Abstract: Acute inhibition of monoamine oxidase B (MAO-B) in the rat does not affect striatal dopamine (DA) metabolism, but chronic MAO-B inhibition with deprenyl has been reported to increase the release of striatal DA, as shown using in vitro techniques. To see whether chronic MAO-B inhibition also causes an increase in DA release in vivo, rats were treated for 21 days with either deprenyl (0.25 mg/kg), TVP-1012 [R(+)-N-propargyl-1-aminoindan mesylate; 0.05 mg/kg), an irreversible inhibitor of MAO-B that is not metabolized to amphetamines, clorgyline (0.2 mg/kg), or saline (all doses once daily by subcutaneous injection). Concentric 4-mm-long microdialysis probes were implanted in the left striatum under pentobarbital/chloral hydrate anesthesia on day 21, and microdialysate DA, 3,4-dihydroxyacetic acid (DOPAC), and 4-hydroxy-3-methoxyphenyl acetic acid (HVA) were determined in the conscious animals on day 22. Baseline levels of DA were as follows: control, 0.34 ± 0.04 (n = 13); deprenyl, 0.88 ± 0.10 (n = 8, p < 0.01); TVP-1012, 0.94 ± 0.20 (n = 7, p < 0.01); clorgyline, 0.90 ± 0.12 (n = 7, p < 0.01) pmol/20 min. Levels of DOPAC and HVA were reduced only in the clorgyline-treated group. The incremental release of DA induced by depolarizing concentration of K+ (100 mM bolus of KCl in perfusate) was significantly greater in clorgyline- and deprenyl-treated rats and elevated (nonsignificantly) in TVP-1012-treated rats. Chronic treatment with the MAO-B inhibitors reduced striatal MAO-B activity by 90%, with 15% (TVP-1012) or 40% (deprenyl) inhibition of MAO-A. Clorgyline inhibited MAO-A by 95%, with 30% inhibition of MAO-B. A single dose of deprenyl (0.25 mg/kg, 24 h before microdialysis) had no significant effect on striatal efflux of DA. The results show that DA metabolism was reduced only by clorgyline, whereas neuronal release of DA was enhanced by both MAO-A and MAO-B inhibitors on chronic administration. The enhanced DA release by chronic MAO-B inhibition does not appear to be dependent on production of amphetamine-like metabolites of the inhibitor. Possible mechanisms for the release-enhancing effect of the MAO-B inhibitors include elevation in levels of endogenous β-phenylethylamine, or an inhibition of DA reuptake, which develops only on chronic administration, because both deprenyl and TVP-1012 have only very weak effects on amine uptake in acute experiments.  相似文献   

15.
Abstract: The effects of intracerebroventricular administration of the 5-hydroxytryptamine (5-HT)1A agonist, 8-hydroxy-2-(di-n-propylamino)tetralin (8-OH-DPAT; 0.1 pmol) on adrenocortical and neurochemical responses to stress were examined in conscious male rats. The following stress paradigms were used: acoustic stimulation (105 dB for 2 min); footshock (0.2 mA, five shocks over 5 min); conditioned fear (animals placed in a footshock chamber for 5 min, 24 h after footshock); restraint (5 min); intraperitoneal (i.p.) injection of recombinant human interleukin-1α (rHu-IL-1α, 20 µg/kg); and injection of cocaine hydrochloride (20 mg/kg, i.p.). As previously shown, 8-OH-DPAT was able to attenuate the adrenocortical response to acoustic stress, conditioned fear, rHu-IL-1α, and cocaine administration. Cocaine decreased 5-hydroxyindoleacetic acid (5-HIAA)/5-HT and dihydroxyphenylacetic acid/dopamine (DOPAC/DA) ratios and norepinephrine (NE) concentration in the prefrontal cortex, hypothalamus, and brainstem in all experiments, and 8-OH-DPAT reversed the changes in DOPAC/DA ratio without affecting 5-HIAA/5-HT ratios or NE content. 8-OH-DPAT alone had no effect on these parameters, although it decreased NE content in the prefrontal cortex in several experiments, and in the brainstem in one experiment. Significant decreases in NE content were observed in some brain regions following some of the stressors, but these changes were not generally affected by 8-OH-DPAT. Increases in the 5-HIAA/5-HT and DOPAC/DA ratios were also observed in some brain sites following some stressors, but these changes were not affected by 8-OH-DPAT except in the case of the increased 5-HIAA/5-HT ratio in the prefrontal cortex following the conditioned fear response. These results indicate that although 8-OH-DPAT is able to decrease plasma corticosterone responses following acoustic stress, conditioned fear, rHu-IL-1α, and cocaine administration, these effects do not appear to be related to an action of the 5-HT1A agonist on biogenic amine metabolism. This observation indicates that the predominant effect of 8-OH-DPAT on adrenocortical responses is mediated at postsynaptic sites not involved in the regulation of cerebral biogenic amine metabolism.  相似文献   

16.
We have obtained and modeled the electrical characteristics of the plasma membrane of Chara internodal cells: intact, without turgor and perfused with and without ATP. The cells were voltage and space-clamped to obtain the I/V (current-voltage) and G/V (conductance-voltage) profiles of the cell membrane. The intact cells yielded similar I/V characteristics with resting p.d.s of −221 ± 12 mV (cytoplasmic clamp, 5 cells) and −217 ± 12 mV (vacuolar clamp, 5 cells). The cut unperfused cells were depolarized at −169 ± 12 mV (7 cells) compared to the vacuole-clamped intact cells. The cells perfused with ATP fell into three groups: hyperpolarized group with resting p.d. −175 ± 12 mV (4 cells) and I/V profile similar to the intact and cut unperfused cells; depolarized group with resting p.d. of −107 ± 12 mV (6 cells) and I/V profiles close to linear; and excited cells with profiles showing a negative conductance region and resting p.d. at −59 ± 12 mV (5 cells). The cells perfused with medium containing no ATP showed upwardly concave I/V characteristics and resting p.d. at −81 ± 12 mV (6 cells). The I/V curves were modeled employing the ``Two-state' model for the H+ pump (Hansen et al., 1981). The inward and outward rectifiers were fitted to exponential functions and combined with a linear background current. The excitation state in perfused cells was modeled by including an inward current, i excit, with p.d.-dependence described by a combination of hyperbolic tangent functions. An inward current, i no-ATP, with a smaller amplitude, but very similar p.d.-dependence was also included in the simulation of the I/V curves from cells without ATP. This approach avoided I/V curve subtraction. The modeling of the total I/V and G/V characteristics provided more information about the parameters of the ``Two-state' pump model, as well as more quantitative understanding of the interaction of the major transport systems in the plasmalemma in generation of the resting potential under a range of circumstances. ATP had little effect on nonpump currents except the excitation current; depolarization profoundly affected the pump characteristics. Received: 23 January/Revised: 10 October 1995  相似文献   

17.
Abstract: The existence of both nitric oxide synthase (NOS) immunoreactive interneurons and amino acid neurotransmitter-mediated nitric oxide (NO) release in the striatum suggests a role for NO in modulating striatal function. To explore the potential interaction between NO and dopaminergic neurotransmission, the NO-releasing agent (±)-S-nitroso-N-acetylpenicillamine (SNAP) was administered locally into the anterior medial striatum of chloral hydrate-anesthetized rats. SNAP, at 0.5, 1, and 2 mM concentrations, elevated striatal extracellular (EC) dopamine (DA) to 200 ± 42, 472 ± 120, and 2,084 ± 496%, respectively, above baseline levels. Perfusion with (±)-penicillamine (PEN, 1 mM), the non-NO-containing carrier component of SNAP, was ineffective, indicating that PEN is not responsible for SNAP-mediated DA release. Additional microdialysis experiments suggest SNAP-mediated DA release is not due to NO-induced neurotoxicity or blockade of the DA transporter. The DA-releasing effect of SNAP was attenuated under calcium-free conditions and abolished in rats pretreated with reserpine (5 mg/kg), implicating a calcium-sensitive vesicular-dependent release process. To determine the mechanism of SNAP-mediated DA release, the guanylyl cyclase (GC) inhibitor LY 83583 (100 µM) was administered 100 min before and during the SNAP pulse. LY 83583 elevated EC DA levels approximately fivefold and potentiated the DA-releasing effect of SNAP to 2,598 ± 551% above basal DA levels. Similar pretreatments with both the noncompetitive N-methyl-d -aspartate (NMDA) antagonist MK-801 (10 µM) and the competitive NMDA-receptor antagonist (±)-3-(carboxypiperazin-4-yl)propyl-1-phosphonic acid [(±)-CPP, 100 µM] blocked SNAP-mediated DA release. SNAP-mediated DA release was also significantly blunted by pretreatment and coperfusion with MgSO4 (10 mM) and 6,7-dinitroquinoxaline-2,3-dione (DNQX, 10 µM) but not (+)-2-amino-3-phosphonopropionic acid (AP-3, 10 µM). These results suggest that NO releases DA via a calcium-sensitive vesicular-dependent process that is independent of GC activation. In addition, NMDA and kainate/(±)-α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor-mediated mechanisms are implicated in NO-induced DA release.  相似文献   

18.
A radioenzymatic assay has been developed for the sensitive determination of plasma catecholamines in perchloric acid extracts using α-methyldopamine as an internal standard. With 25 μl of plasma extract in a total volume of 40 μl the assay gives blank values equivalent to approcximately 2 femtomoles (fmole) for epinephrine (E), norepinephrine (NE), 6 fmole for α-methyldopamine (MeDA) and approximately 15 femtomoles for dopamine (DA). Recoveries of 25 dpm/fmole NE, 40 dpm/fmole E, 56 dpm/fmole DA and 80 dpm/fmole MeDA have been obtained. The assay is linear to at least 1 picomole catecholamine (CA) and shows less than 0.5% crossover between E, NE and DA and a 4.7% crossover of αMeDA into DA. The interassay variability was ± 7% for DA, ± 4% for E and ±3% for NE.  相似文献   

19.
Previous studies have shown that norepinephrine (NE) elicits trans-membrane potential changes in skeletal muscle cells from normal and dystrophic (BIO 14.6) hamsters, with the magnitude of these changes being significantly less in dystrophic cells. To determine if the decreased response of the dystrophic muscle cells reflects a more generalized phenomenon, the present study was designed to evaluate the effects of NE on membrane properties of brown adipocytes. In vivo techniques using glass microelectrodes were similar to those used in the muscle studies. NE injection (2 to 5 μg/kg body wt, i.v.) into anesthetized hamsters was followed by membrane depolarization, the magnitude of which did not significantly differ in the dystrophic and normal adipocytes. For example, upon administration of 5 μg NE/kg body wt, the average depolarization was 14.5 ± 1.3 mV (X ± S.E.) for 20 dystrophic cells and 14.1 ± 1.8 mV for 18 normal cells. The depolarizations following i.v. infusion of isoproterenol and phenylephrine also had similar amplitudes in both normal and dystrophic cells. Despite this lack of difference in plasma membrane responses, NE induced a significantly smaller rise in interscapular brown fat temperature in the dystrophic (0.09°C) than in the normal hamsters (0.26°C) following administration of 5 μg NE/kg body wt. Thus, the decreased responsiveness to NE of dystrophic sarcolemma did not occur with the plasma membrane of brown adipocytes, although brown fat temperature changes in the dystrophic hamsters were decreased in amplitude.  相似文献   

20.
The effect of subcutaneous injections of saline (0.9% NaCl, 10–40 μl/g b. wt) to 5- and 20-day old rats on the concentrations of tyrosine (Tyr) and tryptophan (Trp) in the serum and the brain and on the levels of biogenic amines and their metabolites in the developing brain at 6 h p.i. is described. At day 5 the concentration of Tyr in the blood was decreased (dose-dependent), but the brain concentrations of Tyr and of its amine-metabolites, dopamine (DA), norepinephrine (NE), homovanillic acid (HVA) and dihydroxyphenylacetate (DOPAC) were unaffected. In contrast, in the 20-day old rat, serum Tyr was unaffected by the saline injections, but the Tyr concentration in the brain decreased markedly at the highest saline dose. The concentrations of NE (only at maximum dose) and of DA (independent on the amount of saline injected) were elevated in the brains of saline injected 20-day old rats. The concentrations of Trp and indoles were more affected at day 5 than at day 20: slightly decreased concentration of Trp in the serum but markedly increased concentrations of brain Trp (only at maximum dose), elevated serotonin (5-HT, independent on the amount of saline injected) and 5-hydroxyindoleacetic acid (5-HIAA, at maximum dose) in the brain. If the maximum dose of 40 μl/g body weight was injected to suckling rats repeatedly during the whole suckling period (in 12 h intervals), some effects caused by one single injection of 40 μl/g disappeared (Tyr—depletion in blood or brain, increase in brain NE, DA and Trp), but other additional effects appeared (decreased DA and increased DOPAC, decreased 5-HT and 5-HIAA). The results show that saline injections do cause characteristic, age-dependent alterations of precursor availability as well as of the rate of synthesis and degradation of catecholamine and 5-HT. Repeated treatments have different effects than one single treatment on the precursor availability and the metabolism of monoamines. These alterations must be taken into account if the effects of certain “specific” treatments are compared and discussed in relation to saline “controls”.  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