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 31 毫秒
1.
J L Waddington  T J Crow 《Life sciences》1979,25(15):1307-1314
Rats with unilateral 5,7-DHT lesions, but not 5,6-DHT lesions, showed rotational responses to 5-HTergic drugs (5MeODMT and fenfluramine) that were qualitatively similar to those induced by DAergic drugs (apomorphine and amphetamine) after 6-OHDA lesions. However, 5,7-DHT-lesioned rats also themselves showed rotational responses to DAergic drugs. The merits and limitations of a unilateral 5,7-DHT-lesioned rotating rat model for studying 5-HTergic function are discussed. It is suggested that 5-HT and DA may function in a co-operative manner in the striatum. These findings may be important for the rational pharmacotherapy of Parkinson's disease in which 5-HT as well as DA has been shown to be substantially depleted.  相似文献   

2.
Abstract: When incubated with a hydroxyl radical (HO?)-generating system (ascorbic acid/Fe2+-EDTA/O2/H2O2), 5-hydroxytryptamine (5-HT; serotonin) is rapidly oxidized initially to a mixture of 2,5-, 4,5-, and 5,6-dihydroxytryptamine (DHT). The major reaction product is 2,5-DHT, which at physiological pH exists as its keto tautomer, 5-hydroxy-3-ethylamino-2-oxindole (5-HEO). Rapid autoxidation of 4,5-DHT gives tryptamine-4,5-dione (T-4,5-D), which reacts with the C(3)-centered carbanion of 5-HEO to give 3,3′-bis(2-aminoethyl)-5-hydroxy-[3,7′-bi-1H-indole]-2,4′,5′-3H-trione (7). The latter slowly cyclizes to 3′-(2-aminoethyl)-1′,6′,7′,8′-tetrahydro-5-hydroxyspiro[3H-indole-3,9′-[9H]pyrrolo[2,3-f]quinoline]-2,4′,5′(1H)- trione (9). A minor amount of T-4,5-D dimerizes to give 7,7′-bi-(5-hydroxytryptamine-4-one) (7,7′-D). In the presence of GSH, the reaction of T-4,5-D with 5-HEO is diverted and, in the presence of sufficient concentrations of this tripeptide, completely blocked. This is because GSH preferentially reacts with T-4,5-D to give 7-S-glutathionyltryptamine-4,5-dione (11). The results of this investigation suggest that 5,6-DHT, 5-HEO, 7, and 9 are products unique to the HO?-mediated oxidation of 5-HT. Thus, the observation of other investigators that 5,6-DHT is formed in the brains of rats following a large dose of methamphetamine (MA) suggests that this drug might evoke HO? formation. However, the present in vitro study indicates that 5,6-DHT is a rather minor, unstable product of the HO?-mediated oxidation of 5-HT and suggests that detection of 5-HEO, 7/9, and 11 in rat brain following MA administration could provide additional support for HO? formation. Furthermore, one or more of the intermediates and major products of oxidation of 5-HT by HO? might, in addition to 5,6-DHT, contribute to the MA-induced degeneration of serotonergic neurons.  相似文献   

3.
The receptor specificity for synaptically mediated lateral inhibition in Limulus lateral eye retina was studied by structure-activity correlations of the action of the putative indoleaminergic neurotransmitter, serotonin (5-HT), and its isomers and structural analogs, tryptamine (TRYP), 6-hydroxytryptamine (6HT), 5,6-dihydroxytryptamine (5,6-DHT), 5-hydroxydimethyltryptamine (5-HDMT), and 5-hydroxytryptophan (5-HTP). The 5-HT blockers, lysergic acid diethylamide (LSD), bromo-LSD (BOL), and cinanserin, were also tested. The inhibitory action of the indoleaminergic agonists is highly structure-specific. An hydroxyl group in the 5 position of the indole nucleus, sterically unencumbered by hydroxyls in neighboing positions, is essential. In order of decreasing potency, 5-HT, 5-HDMT, and 5-HTP are active agonists; TRYP, 6-HT, and 5,6-DHT are inactive. Configuration and mobility of the side chains of the active agonists also affect the interaction, and these side-chain characteristics correlate with agonist potency. The receptors for inhibitory action and for transmembranal transport in reuptake are different. Both active agonists and inactive analogs appear to be taken up (Adolph and Ehinger, 1975. Cell Tissue Res. 163:1-14). LSD and BOL have bimodal actions: direct inhibition and agonist blockade. These actions may be mediated via low-specificity presynaptic uptake receptor sites rather than highly specific, postsynaptic, agonist receptor sites.  相似文献   

4.
Rat brain cortex slices preincubated with 3H-5-hydroxytryptamine (3H-5-HT) were superfused with physiological salt solution containing paroxetine, an inhibitor of 5-hydroxytryptamine (5-HT) uptake. The effects of various indolethylamines on the electrically evoked tritium overflow (containing 66.3% unmetabolized 3H-5-HT) were investigated (the percentage of unmetabolized 3H-5-HT was not altered by the indolethylamines or metitepin). 6,7-Dihydroxytryptamine (6,7-DHT) did not affect the stimulation-evoked tritium overflow, whereas the latter was inhibited by the other tryptamine derivatives investigated; when the compounds were compared to each other on the basis of their inhibitory potencies the following rank order was obtained: unlabelled 5-HT > 5-methoxytryptamine > 4-HT > 6-HT > 5,6-DHT > tryptamine > 7-HT > 5,7-DHT. The inhibitory effects of these compounds were antagonized by metitepin. It is concluded that the indolethylamines inhibit the stimulation-evoked 3H-5-HT release by activating the presynaptic 5-HT autoreceptors on the 5-HT neurones of the rat brain cortex. Similarities may exist between these receptors and the postsynaptic 5-HTl binding sites of this brain area.  相似文献   

5.
1. 5-HT (10−8−5 × 10−6 M) relaxed isolated locust foreguts.2. The effects of 5-HT were mimicked by 5,6-DHT, 5-MeOT, tryptamine, 5-MeT, MK212 and methysergide while 5-hydroxyindole, 5-hydroxyindole acetic acid, 5-hydroxytryptophol, NN-DMT, 8-OH DPAT and RU 24969 were without effect.3. Ketanserin (pA2 = 5.65) was a competitive antagonist of the effects of 5-HT, MK212 and methysergide.4. Mianserin (pA2 = 6.3) was also a competititve antagonist of 5-HT but ICS 205-930 had no antagonistic effect on 5-HT-induced relaxation.5. It is concluded that 5-HT relaxes the locust foregut by interacting with 5-HT2-like receptors.  相似文献   

6.
Abstract: Oxygen radicals have been implicated in the neurodegenerative and other neurobiological effects evoked by methamphetamine (MA) in the brain. It has been reported that shortly after a single large subcutaneous dose of MA to the rat, the serotonergic neurotoxin 5,6-dihydroxytryptamine (5,6-DHT) is formed in the cortex and hippocampus. This somewhat controversial finding suggests that MA potentiates formation of the hydroxyl radical (HO?) that oxidizes 5-hydroxytryptamine (5-HT) to 5,6-DHT, which, in turn, mediates the degeneration of serotonergic terminals. A major and more stable product of the in vitro HO?-mediated oxidation of 5-HT is 5-hydroxy-3-ethylamino-2-oxindole (5-HEO). In this investigation, a method based on HPLC with electrochemical detection (HPLC-EC) has been developed that permits measurement of very low levels of 5-HEO in rat brain tissue in the presence of biogenic amine neurotransmitters/metabolites. After intracerebroventricular administration into rat brain, 5-HEO is transformed into a single major, but unknown, metabolite that can be detected by HPLC-EC. One hour after administration of MA (100 mg/kg s.c.) to the rat, massive decrements of 5-HT were observed in all regions of the brain examined (cortex, hippocampus, medulla and pons, midbrain, and striatum). However, 5-HEO, its unidentified metabolite, or 5,6-DHT were not detected as in vivo metabolites of 5-HT. MA administration, in particular to rats pretreated with pargyline, resulted in the formation of low levels of N-acetyl-5-hydroxytryptamine (NAc-5-HT) in all brain regions examined. These results suggest that MA does not potentiate the HO?-mediated oxidation of 5-HT. Furthermore, the rapid MA-induced decrease of 5-HT might not only be related to oxidative deactivation of tryptophan hydroxylase, as demonstrated by other investigators, but also to the inhibition of tetrahydrobiopterin biosynthesis by NAc-5-HT. The massive decrements of 5-HT evoked by MA are accompanied by small or no corresponding increases in 5-hydroxyindole-3-acetic acid (5-HIAA) levels. This is due, in part, to the relatively rapid clearance of 5-HIAA from the brain and monoamine oxidase (MAO) inhibition by MA. However, the loss of 5-HT without corresponding increases in its metabolites point to other mechanisms that might deplete the neurotransmitter, such as oxidation by superoxide radical anion (O2??), a reaction that in vitro does not generate 5-HEO or 5,6-DHT but rather another putative neurotoxin, tryptamine-4,5-dione. One hour after administration, MA evokes large depletions of norepinephrine (NE) throughout the brain but somewhat smaller decrements of dopamine (DA) that are restricted to the nigrostriatal pathway. Furthermore, MA evokes a major shift in the metabolism of both NE and DA from the pathway mediated by MAO to that mediated by catechol-O-methyltransferase. The profound and widespread effects of MA on the noradrenergic system, but more anatomically localized influence on the dopaminergic system, suggests that NE in addition to DA, or unusual metabolites of these neurotransmitters, might play roles in the neurodegenerative effects evoked by this drug.  相似文献   

7.
《Life sciences》1987,41(14):1717-1723
The ergot derivatives, bromocriptine, lisuride and quinpirole (Ly-171555), activators of D-2 receptors, increased striatal acetylcholine (ACh) content by about 40% and induced a 30% inhibition of ACh evoked release from striatal slices, similar to the effects of the dopaminergic agonist apomorphine. These actions were a consequence of dopaminergic activation since they were antagonized by pretreatment with the neuroleptic agent, pimozide. In contrast, pretreatment with L-sulpiride (100 mg/kg), a specific antagonist for the D-2 dopaminergic receptor only, prevented the rise of ACh levels induced by apomorphine or quinpirole but did not interfere with the lisuride- or bromocriptine- induced ACh increases. Similarly, inhibition of the ACh evoked release produced by lisuride (3ωM) was prevented by pimozide (1mg/kg) but not by pretreatment with L-sulpiride. Addition of L-sulpiride (5ωM) to the Krebs solution had no effect on the inhibition of ACh-evoked release induced by lisuride, but a lower concentration (1ωM) antagonized the inhibition induced by quinpirole. Lisuride and bromocriptine responses were both insensitive to sulpiride. These results are discussed in terms of different interaction with the dopaminergic D-2 receptors by the drugs studied.  相似文献   

8.
Previous data (1) have shown that L-DOPA increases the duration of the clonic phase of post-decapitation convulsions (PDC) in mice. It was suggested that this effect is produced by depleting 5-hydroxytryptamine (5-HT) in the inhibitory bulbospinal pathways and thus enhancing reflex activity in the spinal cord. If this were true then L-DOPA administration should not influence clonic PDC in animals whose 5-HT pathways were destroyed. We therefore tested the effects of L-DOPA on mice 3 weeks after pretreatment with the 5-HT neurotoxin, 5,6-dihydroxytryptamine (5, 6-DHT) (50 μg/kg, intracerebroventricularly). All mice were given the peripheral decarboxylase inhibitor, Ro 4-4602. 5,6-DHT halved the brain 5-HT levels and significantly increased the duration of clonic PDC. The administration of L-DOPA (320 mg/kg i.p.) to 5,6 DHT treated mice did not produce any further significant increases in duration. The administration of 5-hydroxytryptophan (5-HTP) (100 mg/kg, i.v.) to 5,6-DHT treated mice, however, increased 5-HT to above control levels and reduced convulsions to control levels. Administration of both 5-HTP and L-DOPA to 5,6-DHT treated mice resulted in 5-HT levels and convulsion times which were also not significantly different from the controls. These data give additional indication that intact 5-HT nerve terminals are necessary for L-DOPA to prolong the duration of clonic PDC.  相似文献   

9.
The present investigation was designed to determine the effect of hallucinogens on the facilitating action of serotonin (5-HT) and norepinephrine (NE) in the facial nucleus. Intravenous administration of d-lysergic acid diethylamide (LSD, 5–10 μg/kg), mescaline (0.5–1.0 mg/kg), or psilocin (0.5–1.0 mg/kg) had no effect by themselves on the glutamate-induced excitation of facial motoneurons. In contrast, the facilitation of facial neuron excitation by iontophoretically applied 5-HT and NE was enhanced 6–10 fold by these hallucinogens. The LSD-enhanced responses to 5-HT and NE continued for at least 4 hours after administration of the hallucinogen. Iontophoretic application of LSD or mescaline (low currents) also markedly potentiated the facilitating effect of 5-HT and NE. Higher currents of LSD (15–40 nA) temporarily antagonized the response to 5-HT. The nonhallucinogen ergot derivatives lisuride and methysergide failed to potentiate the facilitating effects of 5-HT or NE. These observations suggest that hallucinogens potentiate the effect of monoamines on facial motoneurons by increasing the sensitivity of 5-HT and NE receptors. A novel mechanism regarding the psychedelic effects of hallucinogens is discussed.  相似文献   

10.
The neonatal administration of 5,7-dihydroxytryptamine to rats (100 mg kg?1 s.c. on the 1st and 2nd day after birth) resulted in marked reductions in serotoninergic presynaptic markers ([3H]-5-HT synaptosomal uptake, tryptophan hydroxylase activity and endogenous 5-HT content) in various forebrain areas, particularly the cerebral cortex and the hippocampus. In contrast, this treatment produced an increased outgrowth of serotoninergic terminals in the brain stem as judged by the significant increments of these presynaptic markers in this region. Both in the hippocampus and the brain stem, these 5,7-dihydroxytryptamine-induced changes in serotoninergic innervation were associated with a transient increase in 5-HT-sensitive adenylate cyclase activity. No significant alteration of the specific high affinity binding of [3H]-5-HT to synaptosomal membranes from various brain regions was detected in 5,7-dihydroxytryptamine-treated rats for at least the first postnatal month.The chronic blockade of 5-HT receptors by metergoline (5 mg kg?1 day?1 from day 3 to day 22 after birth) altered neither the changes in presynaptic markers nor the evolution of [3H]-5-HT high affinity binding in 5,7-dihydroxytryptamine-treated rats.These findings further illustrate that the high affinity binding sites for [3H]-5-HT do not correspond to postsynaptic 5-HT receptors coupled to adenylate cyclase in the rat brain. Apparently, 5-HT receptors play no role in the increased outgrowth of serotoninergic systems in the brain stem following neonatal 5,7-dihydroxy-tryptamine treatment.  相似文献   

11.
Recent work has shown that intracerebral injections of 5,6-dihydroxytryptamine (5,6-DHT) lead to a fairly selective and long lasting depletion of 5-HT in the rat CNS (BAUMGARTEN, BJORKLUND, LACHENMAYER, NOBIN and STENEVI, 1971; DALY, FUXE and JONSSON, 1973). This effect appears to result from a degeneration of the serotonin-containing neurons (BAUMGARTEN and LACHENMAYER, 1972a). 5,6-DHT does, however, to a lesser extent affect both NA and dopamine (DA) containing nerve terminals (BAUMGARTEN et al., 1971). In an attempt, therefore, to find compounds having a more specific toxic action we have investigated several other hydroxylated tryptamines. In order to obtain information about the differential affinities of these analogues for neuronal uptake sites we have examined their effects on the uptake of [3H]5-HT and (±)-[3H]NA into synaptosomes in homogenates of rat hypothalamus and of [3H]DA uptake into a similar preparation from the rat corpus striatum. It is known that the uptake of these putative transmitters in rat brain homogenates is predominantly into the synaptosome fraction (KANNENGIESSER, HUNT and RAYNAUD, 1973; COYLE and SNYDER, 1969).  相似文献   

12.
Abstract: Spontaneous oxygen consumption by 5,6- and 5,7-DHT (dihydroxytryptamine), related indoleethylamines, and 6-hydroxydopamine and oxygen consumption by these compounds in the presence of rat liver mitochondria were measured by the polarographic oxygen electrode technique. 5,6- and 5,7-DHT react with oxygen at very different rates (2.7 nmol O2/min and 33.4 nmol O2/min, respectively) when incubated in buffer, pH 7.2, at a concentration of 1 mm and with different kínetic characteristics. While the oxidation of 5,7-DHT obeys a reaction of second-order type, the oxidation of 5,6-DHT is more complex and characterized by autocatalytic promotion. Coloured quinoidal oxidation products appeared during the degradation of both indoleamines. Glutathione, ascorbate, dithiothreitol, cysteine, albumin, and superoxide dismutase partially prevented 5,6- and 5,7-DHT from oxidative destruction. Catalase saved oxygen only in the case of 5,6-DHT by recycling of O2 released from near-stoichiometrically formed H2O2 during oxidation of 5,6-DHT: 5,7-DHT did not generate H2O2 in measurable amounts. Oxygen consumption rates of 5,6- and 5,7-DHT were enhanced after addition of rat liver mitochondria to the incubation medium; this resulted in an accelerated formation of quinoidal products. This stimulatory effect on the oxidation rates of both 5,6- and 5,7-DHT was blocked by cyanide, but not rotenone, and was abolished by boiling of the mitochondria fraction. The observed increase in oxygen consumption in the presence of mitochondria was found not to be influenced by monoamine oxidase-dependent deamination of 5,6- and 5,7-DHT. It is postulated that 5,6- and 5,7-DHT are capable of participating in the electron transfer of the mitochondrial respiration chain beyond complex III. Results obtained in determinations of ADP:0 ratios in respiratory control experiments exclude a possible interference of 5,6-DHT, 5,7-DHT, and 6-OH-DA with phosphorylating sites. During the activated state of respiration, no signs of electron transfer inhibition by 5,6- and 5,7-DHT were detectable. A comparison and evaluation of the autoxidation rates of various hydroxylated indoleethylamines, of their affinity to the 5-HT transport sites, and their neurotoxic potency in vivo reveals that interaction of these compounds with oxygen at restricted reaction velocity is a prerequisite for efficient toxicity in monoaminergic neurons following active accumulation in these neurons via the high-affinity uptake systems.  相似文献   

13.
Intraventicular injection of beta-endorphin (beta LPH61?91) in urethane anesthetized male rats led to a dose dependent increase of plasma prolactin levels. Intravenous injection of apomorphine completely abolished the stimulatory effect of beta-endorphin. Animals treated with 6-hydroxydopamine (6-OHDA) and 6-OHDA plus desmethylimipramine showed inhibition of beta-endorphin induced prolactin release. These results suggest that beta-endorphin presynaptically inhibits the activity of dopaminergic neurones, leading to the stimulation of plasma prolactin levels.  相似文献   

14.
[14C]5,6-Dihydroxytryptamine ([14C] 5,6-DHT) and [14C]5,7-dihydroxytryptamine ([14C]5,7-DHT) were deaminated to toluene-isoamylalcohol extractable products when incubated with homogenates of rat hypothalamus or pons-medulla oblongata. [14C]5,6-Dihydroxyindole acetic acid ([14C]5.6-DHIAA) and [14C]5,7-dihydroxyindole acetic acid ([14C]5,7-DHIAA) were detected as MAO metabolites by TLC besides non-identified components. The conversion of [14C]5,6-DHT and [14C]5,7-DHT obeyed, at least initially, Michaelis-Menten kinetics (Km 5,7-DHT: 0.5 × 10?3M; Km 5,6-DHT: 1.25 × 10?3M). Inhibition of the reaction by the MAO A inhibitor, clorgyline, resulted in a typical double sigmoidal inhibition curve indicating that both amines are metabolized by both types of MAO (A and B). In deprenyl inhibition studies, however, 5,7- and 5,6-DHT seemed to be preferred substrates of MAO A. Incubation of rat brain homogenates with [14C]5,6-DHT and [14C]5,7-DHT or with the MAO metabolites [14C]5,6-DHIAA and [14C]5,7-DHIAA caused a time-dependent break-down of the dihydroxylated indole compounds with subsequent binding of radioactivity to perchloric acid insoluble tissue components. 5,6-DHT inactivated MAO in rat brain homogenates parallel to its decomposition and extensive protein binding. The inactivation of MAO by 5,6-DHT and the extensive binding of radioactivity to protein were antagonized by dithiothreitol (DTT), glutathione (GSH) and L-ascorbic acid. Reduction of [O2] in the incubation medium slightly attenuated the inactivation of MAO by 5,6-DHT. Catalase or superoxide dismutase failed to prevent MAO from being inactivated by 5,6-DHT. The results suggest that oxidation products of 5,6-DHT, e.g. its corresponding o-quinone, are involved in the inactivation of MAO in vitro and mainly responsible for the binding of radioactivity to brain proteins in vitro. Similar mechanisms may also be operative in the in vivo neurotoxicity of 5,6-DHT. The lack of inactivation of MAO by 5,7-DHT in vitro correlated with a low degree of radioactivity binding (from [14C]5,7-DHT) to homogenate protein pellets; the binding to proteins was barely influenced by GSH, cysteine, DTT and l -ascorbic acid. These latter findings do not provide a plausible explanation for the mechanism(s) involved in the well known in vivo neurotoxicity of 5,7-DHT.  相似文献   

15.
1. Dialysed serotonergic neurons were identified, isolated from the ganglia of 5,6-dihydroxytryptamine (5,6-DHT) treated snail, Helix pomatia L. Twenty-four to 40 days after injection of 5,6-DHT into the animal, serotonergic neurons show a specific brown pigmentation, which stays there for several weeks. After protease digestion (0.5–1.0 mg/ml for 10–12 min) the labelled neurons can be easily separated. This method ensures the reliable identification of serotonergic neurons for intracellular dialysis.2. We showed that isolated serotonergic neurons maintain their membrane characteristics, and ion-currents can be registered under voltage clamp, just as from neurons of untreated animals. The threshold concentration of serotonin (10 −7 M) and the survival time of pigment labelled dissociated cells were the same as for the control cells.3. Following 5-HT application, the voltage activated Ca-currents were either increased or decreased, depending on the neuron used.4. The different responses are probably caused by different receptors on the cell membrane or by the presence of different types of Ca-channels.5. The deactivation time constant of the Ca-current, calculated from the tail current, was also altered in the pigment labelled neuron following serotonin treatment.  相似文献   

16.
6-Ethyl-9-oxaergoline (EOE) and its enantiomers were compared with apomorphine in a number of tests designed to measure dopamine (DA) agonist activity within the central nervous system. In rats, the tests were: interaction with DA receptors labeled with 3H-apomorphine or 3H-spiroperidol; the effects on DA synthesis as assessed by the
-butyrolactone procedure; turning in 6-OHDA lesioned animals; stereotypy; and, slowing of DA cell firing rates. In the mouse, locomotor activity, hypothermia and postural asymmetry in Caudectomized animals were studied. Emesis in the beagle was also examined. The (-)-enantiomer of EOE was more potent than either the (+)-enantiomer or the racemate in all tests. With the exception of inducing stereotypy and the displacement of 3H-apomorphine from rat striatal membranes, (-)-EOE was equi- or more potent than apomorphine in all test procedures. (-)-EOE was effective following oral administration and exhibited a longer duration of action than apomorphine. The results indicate EOE is a potent DA agonist.  相似文献   

17.
To study the early effects of neonatal 5,7-dihydroxytryptamine lesions on 5-hydroxytryptamine1A (5-HT1A) receptors, we measured regional [3H]8-OH-DPAT-labeled 5-HT1A sites in binding assays and compared them to our previous studies of [3H]paroxetine-labeled 5-HT transporter sites during the first month in the same rats. While there were significant time- and dose-dependent effects of 5,7-DHT on 5-HT transporter sites, there were no significant changes in 5-HT1A sites in cortex, hippocampus, diencephalon, brainstem, cerebellum, or spinal cord. 5,7-DHT lesions also did not alter the Ki of Gpp(NH)p at brainstem 5-HT1A sites or the Ki of 5-HT in cortex or brainstem in the presence or absence of GTPS or Gpp(NH)p. There were significant regional differences between the density of 5-HT1A sites and 5-HT transporter sites. The ontogeny of brainstem 5-HT1A sites was a pattern of increases until three weeks postnatal, and 5,7-DHT lesions did not alter the ontogeny of 5-HT1A sites. These data suggest differential plasticity of 5-HT1A and 5-HT transporter binding sites during the first month after neonatal 5,7-DHT lesions.  相似文献   

18.
Copper sulfate (Cu2+, 10?6M to 10?5M) stimulated the rate of autoxidation of the catecholamine neurotoxin 6-hydroxydopamine (6-OHDA) as judged both by absorbancy measurements at 490 nm and oxygen consumption. This stimulation of 6-OHDA autoxidation by Cu2+ was prevented by the copper chelating agent EDTA. In other experiments, Cu2+ stimulated the rates of autoxidation of dopamine and norepinephrine as well as the neurotoxic agents 6-aminodopamine, 5,6-dihydroxytryptamine and 5,7-dihydroxytryptamine. It is suggested that intraneuronal levels of copper, by virtue of controlling the rates of autoxidation reactions, might be important in vivo.  相似文献   

19.
A biochemical study of the endogenous levels of serotonin (5-HT), noradrenaline (NA) and the activity of choline acetyltransferase (CAT) was carried out in the intestinal tract of the rat. High levels of 5-HT and NA were detected in the caecum and the colon. These anatomical regions also presented the highest activity of CAT. Similar activities of CAT were detected, after dissection, in the mucosa and the muscular layers containing the enteric plexuses. During the day-night cycle, 5-HT and NA amounts showed significant variations as a function of time. Treatment with pargyline (75 mg kg?1), a monoamine oxidase inhibitor, resulted in an increase in 5-HT content with parallel modifications in CAT activity. In spite of an important decrease in 5-HT endogenous level in the caecum of rats pretreated with parachlorophenylalanine (300 mg kg?1), no significant change in CAT activity was detected whatever was the duration of the treatment. α-Methylparatyrosine (100 mg kg?1), known to block the synthesis of NA, did not affect the CAT activity in the caecum.  相似文献   

20.
Summary In an attempt to determine the conditions which permit central 5-HT neurons to respond to a chemical injury of their axons by sprouting and regeneration, the pattern and time-course of recovery of 5-HT concentrations and regrowth of bulbospinal 5-HT axons were evaluated in rats subjected to intraventricular treatment with either 75 g 5,6- or 150 g 5,7-DHT. While 5,6-DHT treatment is followed by a significant recovery of 5-HT concentrations in the telodiencephalon, brainstem and upper part of the spinal cord within 3 months, there is no significant restoration of the severely depleted 5-HT levels in the telodiencephalon and spinal cord, and only limited recovery in 5-HT content of the brainstem preparation after 5,7-DHT.These differences conform to the observation of widespread and effective regrowth and regeneration of the bulbospinal 5-HT neurons in the 5,6-DHT treated lower brainstem and upper spinal cord but restricted and localized sprouting efforts in the 5,7-DHT treated lower medulla oblongata. This could be explained by a cell body near lesion of the non-terminal indoleamine axons by 5,7-DHT which results in a late retrograde, irreversible degeneration of most of the indoleamine pericarya from group B1 and many of group B3.It is concluded that the preservation of a critical length of the main axon and part of its collaterals is necessary for the neuron's survival, and that the individual pattern of the neuropil architecture of brain centres which are invaded by the axonal sprouts may significantly influence their growth characteristics and thus either favour or impede their chance to reestablish connections with their original effector. Aberrant, localized, intense sprouting of drug-damaged axons may in itself reflect the need of the neuron—deprived of most of its axonal tree—to reestablish its original total axonal length by multiple branching.Supported by grants from the Deutsche Forschungsgemeinschaft. The authors are indebted to Rolf Franck for his technical assistance.Supported by grants from the Swedish Medical Research Council (No. 04 X-3874 and 04 X-56).  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