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 0 毫秒
1.
Kindling-induced potentiation in the piriform cortex   总被引:2,自引:0,他引:2  
At intensities sufficient to induce epileptiform afterdischarges, repeated electrical stimulation of limbic structures can lead to the development of permanent increases in the strength of the epileptiform response (kindling). Field potentials evoked by pulse stimulation are also increased in amplitude in a number of forebrain pathways following kindling. This kindling-induced potentiation effect is similar in many respects to the 'long-term potentiation' (LTP) effect which is produced by non-epileptogenic stimulation. There are, however, some interesting differences. For example, kindling-induced potentiation can far outlast LTP. In these experiments, we attempted to determine the longevity of the kindling-induced potentiation of the response evoked in the piriform cortex by olfactory bulb stimulation, following olfactory bulb kindling. This system was targeted because both the olfactory bulb and the piriform cortex are highly reactive kindling sites. In addition, we used the paired pulse technique to monitor facilitation and inhibition in this system. Kindling was found to induce a potentiation in the piriform field potential that lasted for at least 3 months (the period of the experiment) with little or no decay. Kindling also produced a decrease in paired pulse facilitation. In some animals the net facilitation was changed to a net depression. These results are consistent with the interpretation that kindling produces an increase in recurrent inhibition in the piriform cortex. The paired pulse measures, however, returned to near baseline levels over the 3-month test period.  相似文献   

2.
We performed whole-cell recordings of layer III non-pyramidal neurons in the piriform cortex of Sprague–Dawley rats. For comparison purposes, recordings were made from deep pyramidal cells, which are also present in layer III. These two cell types could be distinguished both anatomically and physiologically. Anatomically, the layer III non-pyramidal neuron displayed smooth beady dendrites, while deep pyramidal cells showed thicker dendrites with spines. The dendrites of the layer III non-pyramidal neuron also tended to be restricted to layer III while deep pyramidal cells had long apical dendrites that spanned layers I and II. Although the resting membrane potentials of both cell types were very similar, significant differences were noted in other physiological measures. Layer III non-pyramidal neurons typically displayed higher input resistances, faster time constants, smaller spike amplitudes, shorter spike widths, and higher spike thresholds. In addition, layer III non-pyramidal neurons were able to spike at much higher rates when stimulated with the same level of threshold normalized current injection. The most dramatic differences in physiology were seen in the pattern of spiking in response to increasing levels of positive constant current pulses. Layer III non-pyramidal neurons showed qualitatively different responses at low and high levels of stimulation. At low levels, spikes occurred with long latency and the firing frequency increased throughout the duration of the current pulse. At high levels, non-pyramidal neurons started spiking with short latency, followed by a decrease in firing frequency, which in turn was followed by an increase in firing frequency. Deep pyramidal neurons differed dramatically from this pattern, displaying a qualitatively similar response at all levels of current injection. This response was characterized by short latency spikes and spike adaptation for the duration of the current pulse.  相似文献   

3.
Phosphate-activated glutaminase (PAG), which catalyses conversion of glutamine to glutamate, is a potential marker for glutamatergic, and possibly GABA, neurons in the central nervous system. A polyclonal antibody, raised in rabbits against rat brain PAG, was applied to postmortem human brain tissue to reveal the distribution of PAG in the cerebral cortex. PAG immunoreactivity was observed in pyramidal and non-pyramidal neurons but not in glial cells. In the neocortex, large to medium-sized pyramidal neurons in layers III and V were stained most intensely, while the majority of smaller pyramidal cells were labeled either lightly or moderately. Such modified pyramids as the giant Betz cells, the large pyramidal cells of Meynert, and the solitary cells of Ramón y Cajal were also stained intensely. Fusiform cells in layer VI showed moderate to intense labeling. A number of cortical non-pyramidal neurons of various sizes stained moderately to intensely. These included large basket cells which were identified by their characteristic morphology and size in primary cortical areas. Pyramidal cells in the hippocampal formation as well as basket cells of the stratum oriens stained moderately to intensely. Since pyramidal cells are believed to be glutamatergic and large basket cells GABAergic, these results suggest that PAG plays a role in generating not only transmitter glutamate, but also GABA precursor glutamate.  相似文献   

4.
Basket cells, defined by axons that preferentially contact cell bodies, were studied in rat piriform (olfactory) cortex with antisera to gamma-aminobutyric acid (GABA)ergic markers (GABA, glutamate decarboxylase) and to peptides and calcium binding proteins that are expressed by basket cells. Detailed visualization of dendritic and axonal arbors was obtained by silver-gold enhancement of staining for vasoactive intestinal peptide (VIP), cholecystokinin (CCK), parvalbumin, and calbindin. Neuronal features were placed into five categories: soma-dendritic and axonal morphologies, laminar distributions of dendritic and axonal processes, and molecular phenotype. Although comparatively few forms were distinguished within each category, a highly varied co-expression of features from different categories produced a "combinatorial explosion" in the characteristics of individual neurons. Findings of particular functional interest include: dendritic distributions suggesting that somatic inhibition is mediated by feedforward as well as feedback pathways, axonal variations suggesting a differential shaping of the temporal aspects of somatic inhibition from different basket cells, evidence that different principal cell populations receive input from different combinations of basket cells, and a close association between axonal morphology and molecular phenotype. A finding of practical importance is that light microscopic measurements of boutons were diagnostic for the molecular phenotype and certain morphological attributes of basket cells. It is argued that the diversity in basket cell form in the piriform cortex, as in other areas of the cerebral cortex, reflects requirements for large numbers of specifically tailored inhibitory processes for optimal operation that cannot be met by a small number of rigidly defined neuronal populations.  相似文献   

5.
Projections from the amygdala to the piriform cortex are proposed to provide a pathway via which the emotional system can modulate the processing of olfactory information as well as mediate the spread of seizure activity in epilepsy. To understand the details of the distribution and topography of these projections, we injected the anterograde tracer Phaseolus vulgaris-leucoagglutinin into different nuclear divisions of the amygdaloid complex in 101 rats and analyzed the distribution and density of projections in immunohistochemically processed preparations. The heaviest projections from the amygdala to the piriform cortex originated in the medial division of the lateral nucleus, the periamygdaloid and sulcal subfields of the periamygdaloid cortex, and the posterior cortical nucleus. The heaviest terminal labeling was observed in layers Ib and III of the medial aspect of the posterior piriform cortex. Lighter projections to the posterior piriform cortex originated in the dorsolateral division of the lateral nucleus, the magnocellular and parvicellular divisions of the basal and accessory basal nuclei, and the anterior cortical nucleus. The projections to the anterior piriform cortex were light and originated in the dorsolateral and medial divisions of the lateral nucleus, the magnocellular division of the basal and accessory basal nuclei, the anterior and posterior cortical nuclei, and the periamygdaloid subfield of the periamygdaloid cortex. The results indicate that only selective amygdaloid nuclei or their subdivisions project to the piriform cortex. In addition, substantial projections from several amygdaloid nuclei converge in the medial aspect of the posterior piriform cortex. Via these projections, the amygdaloid complex can modulate the processing of olfactory information in the piriform cortex. In pathologic conditions such as epilepsy, these connections might provide pathways for the spread of seizure activity from the amygdala to extra-amygdaloid regions.  相似文献   

6.
Previous studies have suggested that the patterns of innervation and high interconnectivity of the piriform cortex (PC) provide for strong olfactory hippocampal memory; however, these same attributes may create high seizurogenic tendencies. Thus, understanding this wiring is important from a physiological and pathophysiological perspective. Distinct interneurons expressing differing calcium binding proteins (CBPs), parvalbumin (PV), calbindin (CB), and calretinin (CR), have been shown to exist in PC. However, a comprehensive examination of the distribution and innervation patterns of these neurons has not been done. Thus the purpose of this study was to combine the analysis of the CBP cell localization with analysis of their innervation patterns. Each type was differentially localized in the three layers of the PC. Only CR‐positive neurons were found in layer 1. PV and CB are coexpressed in layers 2–3, most expressing both PV and CB. A morphological estimate of the dendritic extent for each subtype showed that PV and PV/CB cells demonstrated equally wide, horizontal and vertical arborizations, whereas CB cells had wide horizontal and restricted vertical arborizations. CR cells had restricted horizontal and very long vertical arborizations. Postsynaptic morphological targeting was also found to be specific, namely, PV+ and PV/CB+ nerve terminals (NTs) innervate perisomatic regions of principal cells. CR+ NTs innervate only dendrites of principal cells, and CB+ NTs innervate both somata and dendrites of principal cells. These data show highly complex innervation patterns for all of the CBP interneurons of the PC and form a basis for further studies in the plasticity of this region. J. Comp. Neurol. 518:1570–1588, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

7.
Learning-related cellular modifications were studied in the rat piriform cortex. Water-deprived rats were divided to three groups: ‘trained’ rats were trained in a four-arm maze to discriminate positive cues in pairs of odours, ‘control’ rats were ‘pseudo-trained’ by random water rewarding, and ‘naive’ rats were water-deprived only. In one experimental paradigm, the trained group was exposed to extensive training with rats learning to discriminate between 35 and 50 pairs of odours. Piriform cortex pyramidal neurons from ‘trained’, ‘control’ and ‘naive’ rats did not differ in their passive membrane properties and single spike characteristics. However, the after-hyperpolarizations (AHPs) that follow six-spike trains were reduced after ‘extensive training’ by 43% and 36% compared with ‘control’ and ‘naive’, respectively. This effect was not observed in the piriform cortex of another group of rats, in which hyperexcitability was induced by chemical kindling. In another experimental paradigm rats were trained only until they demonstrated ‘rule learning’, usually after discriminating between one and two pairs of odours (‘mild training’). In this experiment, a smaller, yet significant, reduction (20%) in AHPs was observed. AHP reduction was apparent in most of the sampled neurons. AHP remained reduced up to 3 days after the last training session. 5 days or more after the last training session, AHP amplitude recovered to pre-training value and did not differ between ‘trained’ rats and the others. Accordingly, training suspension for 5 days or more resulted in slower learning of novel odours. We suggest that increased neuronal excitability, manifested as reduced AHP, is related to the ability of the cortical network to enter a ‘learning mode’ which creates favourable conditions for enhanced learning capability.  相似文献   

8.
Potassium channels of the Kir2 family are widely expressed in neurons and glia, where they form strong inwardly rectifying channels. Existing functional hypotheses for these channels in neurons are based on the weak outward conductance, whereas the leading hypothesis for glia, that they promote potassium spatial buffering, is based on inward conductance. Although the spatial buffering hypothesis has been confirmed for Müller glia in retina, many aspects of Kir2 channels that will be required for understanding their functional roles in neurons and other forms of glia have received little or no study. Particularly striking is the paucity of data regarding their cellular and subcellular localization. We address this gap for Kir2.1-containing channels by using light and electron microscopic immunocytochemistry. The analysis was of piriform cortex, a highly epileptogenic area of cerebral cortex, where pyramidal cells have K(+)-selective strong inward rectification like that observed in Müller cells, where Kir2.1 is the dominant Kir2 subunit. Pyramidal cells in adult piriform cortex also lack I(h), the mixed Na(+)-K(+) current that mediates a slower form of strong inward rectification in large pyramidal cells in neocortex and hippocampus. The experiments demonstrated surface expression of Kir2.1-containing channels in astrocytes and in multiple populations of pyramidal and nonpyramidal cells. Findings for astrocytes were not consistent with predictions for K(+) spatial buffering over substantial distance. However, findings for pyramidal cells suggest that they could be a conduit for spatially buffering K(+) when it is highly elevated during seizure.  相似文献   

9.
We studied the contribution of GABAergic (gamma-aminobutyric acid) neurotransmission to epileptiform activity using the horizontal hippocampal rat brain slice. Seizure-like (ictal) activity was evoked in the CA1 area by applying high-frequency trains (80 Hz for 2 s) to the Schaffer collaterals. Whole-cell recordings from stratum oriens-alveus interneurons revealed burst firing with superimposed high-frequency spiking which was synchronous with field events and pyramidal cell firing during ictal activity. On the other hand, interictal interneuronal bursts were synchronous with large-amplitude inhibitory postsynaptic potentials (IPSPs) in pyramidal cells. Excitatory and inhibitory postsynaptic potentials were simultaneously received by pyramidal neurons during the ictal afterdischarge, and were synchronous with interneuronal bursting and field potential ictal events. The GABAA receptor antagonist bicuculline greatly reduced the duration of the ictal activity in the CA1 layer, and evoked rhythmic interictal synchronous bursting of interneurons and pyramidal cells. With intact GABAergic transmission, interictal field potential events were synchronous with large amplitude IPSPs (9.8 +/- 2.4 mV) in CA1 pyramidal cells, and with interneuronal bursting. Simultaneous dual recordings revealed synchronous IPSPs received by widely separated pyramidal neurons during ictal and interictal periods, indicative of widespread interneuronal firing synchrony throughout the hippocampus. CA3 pyramidal neurons fired in synchrony with interictal field potential events recorded in the CA1 layer, and glutamate receptor antagonists abolished interictal interneuronal firing and synchronous large amplitude IPSPs received by CA1 pyramidal cells. These observations provide evidence that the interneuronal network may be entrained in hyperexcitable states by GABAergic and glutamatergic mechanisms.  相似文献   

10.
The laminar distribution and structure of the supragranular cells projecting from primary auditory cortex (AI) to the second auditory cortex (AII) in the cat were studied with horseradish peroxidase. Injections in AII retrogradely labeled somata in ipsilateral cortical layers I-VI of AI. A bimodal laminar disposition was found, with approximately 40% of the labeled cells in layer III, 25% in layer V, and 10-15% each in layers II, IV, and VI; only a few cells were found in layer I. The labeled cells were scattered in small aggregates between which unlabeled neurons were interspersed. There was some, though not a strict, topographical distribution of the labeled cells according to the locus of the injection in AII. Injections in the caudal part of AII labeled cells in more rostral AI, while rostral AII injections labeled cells in more caudal AI. Ventral AII injections labeled more ventrally located AI cells, while more dorsal AII injections labeled more dorsally situated AI cells. AII injections also labeled cells in other auditory cortex subdivisions, including the posterior ectosylvian, ventroposterior, temporal, and dorsal auditory zone/suprasylvian fringe cortical areas, and in some non-auditory cortical areas. In layers II and III, both pyramidal and non-pyramidal cells were labeled. More pyramidal cells were labeled in layer III than layer II (80% vs. 62%), and the proportion of non-pyramidal cells in layer II was more than twice that in layer IV (27% vs. 12%). The types of labeled cells were distinguished from one another on the basis of size, somatic and dendritic shape, and laminar distribution. The profiles of labeled cells in these experiments were compared to, and correlated with, those in Golgi-impregnated material. In layer II, the classes of corticocortical projecting cells consisted of small and medium-sized pyramidal, bipolar, and multipolar cells. Those in layer III included small, medium-sized, and large pyramidal neurons, and bipolar and multipolar cells. The average somatic area of the labeled cells did not differ significantly from that of the unlabeled cells, and both were about equal in somatic size to neurons accumulating tritiated gamma-aminobutyric acid in layers II and III. These findings suggest that there is convergent, ipsilateral input onto AII from every layer in AI, and from other cortical auditory and non-auditory areas. A morphologically heterogeneous population of cells in AI contributes to these projections. Diversity in the cytological origins of corticocortical projections implies functional differences between layers II and III since the latter also projects commissural, while layer II in the cat, does not.  相似文献   

11.
Multipotent neural precursors have been suggested to exist in many parts of the adult mammalian brain. In the present study, we characterized the neurogenic potential in the piriform cortex of adult rats. Proliferation rates as detected by 5'-bromodeoxyuridine-labeling proved to be low when compared with the major neurogenic brain regions (i.e. the hippocampus and the subventricular zone). 5'-Bromodeoxyuridine/NeuN-labeling in accordance with doublecortin, polysialylated neural cell adhesion molecule, and TUC-4-labeling indicated that neuronal differentiation of newborn cells occurs predominantly in layer II of the piriform cortex. Many of the cells exhibited a pyramidal cell morphology. The lack of 5'-bromodeoxyuridine/NeuN-labeled cells 12 weeks after 5'-bromodeoxyuridine administration argued against long-term survival of newborn neurons in the piriform cortex.  相似文献   

12.
In the present study we analyzed aquaporin-4 (AQP4) immunoreactivity in the piriform cortex (PC) and the hippocampus of pilocarpine-induced rat epilepsy model to elucidate the roles of AQP4 in brain edema following status epilepticus (SE). In non-SE-induced animals, AQP4 immunoreactivity was diffusely detected in the PC and the hippocampus. AQP4 immunoreactivity was mainly observed in the endfeet of astrocytes. Following SE the AQP4-deleted area was clearly detected in the PC, not in the hippocampus. Decreases in dystrophin and α-syntrophin immunoreactivities were followed by reduction in AQP4 immunoreactivity. These alterations were accompanied by the development of vasogenic edema and the astroglial loss in the PC. In addition, acetazolamide (an AQP4 inhibitor) treatment exacerbated vasogenic edema and astroglial loss both in the PC and in the hippocampus. These findings suggest that SE may induce impairments of astroglial AQP4 functions via disruption of the dystrophin/α-syntrophin complex that worsen vasogenic edema. Subsequently, vasogenic edema results in extensive astroglial loss that may aggravate vasogenic edema.  相似文献   

13.
Several lines of evidence indicate a substantial contribution of kainate receptors to temporal lobe seizures. The activation of kainate receptors located on hippocampal inhibitory interneurons was shown to reduce GABA release. A reduced GABA release secondary to kainate receptor activation could contribute to an enhanced seizure susceptibility. As the dentate gyrus serves a pivotal gating function in the spread of limbic seizures, we tested the role of kainate receptors in the regulation of GABA release in the dentate gyrus of control and kindled animals. Application of glutamate (100 micro m) in the presence of the NMDA receptor antagonist d-APV and the AMPA receptor antagonist, SYM 2206 caused a slight depression of evoked monosynaptic inhibitory postsynaptic currents (IPSCs) in control, but a substantial decrease in kindled dentate granule cells. The observation that kainate receptor activation altered paired-pulse depression and reduced the frequency of TTX-insensitive miniature IPSCs without affecting their amplitude is consistent with a presynaptic action on the inhibitory terminal to reduce GABA release. In kindled preparations, neither glutamate (100 micro m) nor kainate (10 micro m) applied in a concentration known to depolarize hippocampal interneurons led to an increase of the TTX-sensitive spontaneous IPSC frequency nor to changes of the postsynaptic membrane properties. Consistently, the inhibitory effect on evoked IPSCs was not affected by the presence of the GABAB receptor antagonist, CGP55845A, thus excluding a depression by an enhanced release of GABA acting on presynaptic GABAB receptors. The enhanced inhibition of GABA release following presynaptic kainate receptor activation favours a use-dependent hyperexcitability in the epileptic dentate gyrus.  相似文献   

14.
Local inhibition by GABA-releasing neurons is important for the operation of sensory cortices, but the details of these inhibitory circuits remain unclear. We addressed this question in the olfactory system by making targeted recordings from identified classes of inhibitory and glutamatergic neurons in the piriform cortex (PC) of mice. First, we looked for feedforward synaptic inhibition provided by interneurons located in the outermost layer of the PC, layer Ia, which is the unique recipient of afferent fibers from the olfactory bulb. We found two types of feedforward inhibition: a fast-rising, spatially restricted kind that was generated by horizontal cells, and a slow-rising, more diffuse kind generated by neurogliaform cells. Both cell types targeted the distal apical dendrites of layer II principal neurons. Next, we studied feedback synaptic inhibition in isolation by making a tissue cut across layer I to selectively remove feedforward inhibitory connections. We identified a powerful type of feedback inhibition of layer II neurons, mostly generated by soma-targeting fast-spiking multipolar cells in layer III, which in turn were driven by feedforward excitation from layer II semilunar cells. Dynamic clamp simulation of feedback inhibition revealed differential effects of this inhibition on the two main types of layer II principal neurons. Thus, our results articulate the connectivity and functions of two important classes of inhibitory microcircuits in the PC. Feedforward and feedback inhibition generated by these circuits is likely to be required for the operation of this sensory paleocortex during the processing of olfactory information.  相似文献   

15.
16.
The types of layer III neurons in cat primary auditory cortex (AI) projecting to the contralateral AI were studied with horseradish peroxidase or horseradish peroxidase conjugated to wheat germ agglutinin. Injections between the anterior and posterior ectosylvian sulci retrogradely labeled both pyramidal and non-pyramidal somata in contralateral cortical layers III, V, and VI in AI, and in the ventral nucleus of the ipsilateral medial geniculate body. Three-quarters (72%) of the retrogradely labeled cells were found in layer III and one-quarter (28%) lay in layers V and VI. Every part of AI was innervated by commissural neurons. The topographical distribution of the labeled cells varied systematically. Injections in the caudal part of AI labeled cells in the caudal part of the opposite AI, while more rostral injections labeled cells in the contralateral, rostral AI. Injections covering the rostro-caudal extent of AI labeled cells throughout the opposite AI. Each part of AI thus projects most strongly to a contralateral, homotypic area, and less strongly to other, adjacent sectors of AI. The types of labeled cells were distinguished from one another on the basis of size, somatic and dendritic morphology, laminar distribution, and nuclear membrane morphology. Their somatodendritic profiles were compared to, and correlated with, those in Golgi-impregnated material from adult animals. Among the pyramidal cells of origin were small, medium-sized, and large neurons, and star pyramidal cells. The non-pyramidal cells of origin included bipolar and multipolar cells. Thus, at least six of the 12 kinds of neurons, as defined by morphological methods, participate in the interhemispheric pathway. Pyramidal cells comprised 65% of the cells of origin, 14% of the labeled cells in layer III were non-pyramidal, and 21% of the neurons could not be classified. It is unknown if these different types of commissural neurons have the same laminar or cytological targets in AI, or if they represent more than one functional or parallel pathway within AI. In any case, cytologically diverse layer III neurons contribute to the commissural system.  相似文献   

17.
Kim DS  Kwak SE  Kim JE  Won MH  Suh JG  Oh YS  Kang TC 《Brain research》2005,1065(1-2):125-131
In the present study, to elucidate the effect of altered P(2)X receptor transmission on GABA(A) receptor expression and its transmission, we studied the morphological and electrophysiological responses of GABA(A) receptor in the gerbil hippocampus following P(2)X receptor antagonist/agonist treatment. Suramin or pyridoxal phosphate-6-azophenyl-2',4'-disulfonic acid (PPADS) treatment did not affect GABA(A) receptor immunoreactivities and paired-pulse responses in the gerbil hippocampus. In addition, ATP treatment did not significantly affect population spike amplitude ratios and EPSP slope ratios in the gerbil dentate gyrus. Co-application, but not pretreatment, of PPADS or suramin enhanced the effect of muscimol on paired-pulse inhibition in the dentate gyrus. In contrast, co-application of ATP reduced the effect of muscimol in the dentate gyrus. These findings indicate that the blockade of P(2)X receptor did not affect GABA(A) receptor immunoreactivities, and P(2)X receptor may modulate GABA(A) receptor-mediated inhibition when in co-activation with GABA(A) receptor. Therefore, our findings suggest that the relationship between GABA(A) receptor and P(2)X receptor may not be reciprocal, although GABA(A) receptor activity affects P(2)X receptor functionality and its expression.  相似文献   

18.
Several studies have shown that the piriform cortex is involved in learning processes and pyramidal cell activity does not only encode the odour quality but is also related to contextual information about past experience and future action. To study how odour-specific patterns in neuronal activity are established we used an odour discrimination go/no go task with water reinforcement for analysing extracellular single cell activity in anterior piriform cortex in freely moving rats. During conditioning single cells responded to different task events. Of the cells 52% participate in odour sampling and 87% were involved in odour discrimination. More than half of the responses to odours were inhibitory responses. Seventeen percent changed their activity for nose-poke only. The activity of 33% was related to reinforcement. Once established the pattern of reaction to the odour was preserved for several days. It is suggested that the anterior part of the piriform cortex is not involved in odour coding only. However, learning-related plasticity was not observed in this area.  相似文献   

19.
The entorhinal cortex receives input from many cortical areas and mediates the flow of information between these sites and the hippocampal formation. Long-term synaptic plasticity in cortical efferents to the entorhinal cortex may contribute to the transmission of neural activity to the hippocampus, as well as the storage of information, but little is known about plasticity in these pathways. We describe here the use of evoked field potential recordings from chronically implanted electrodes in the rat entorhinal cortex to investigate synaptic plasticity in the large piriform (olfactory) cortex projection to the superficial layers of the entorhinal cortex. Both kindling-induced potentiation and long-term potentiation (LTP) were tested. In addition, we attempted to modulate LTP induction by the co-induction of frequency potentiation and by the co-activation of the medial septum. Epileptogenic kindling stimulations of the piriform cortex (1-s, 60-Hz trains 3 times/day for 5 days) were found to result in a reliable potentiation of field responses evoked by piriform cortex test pulses. Non-epileptogenic tetanization of the piriform cortex with 400-Hz 16-pulse trains reliably resulted in LTP effects. These effects could be augmented by embedding brief LTP induction stimuli within 11-pulse, 15-Hz trains that alone produce only frequency potentiation. Co-activating the medial septum with 10-Hz trains, just prior to tetanization of the piriform cortex, augmented LTP of piriform cortex inputs to the entorhinal cortex in an input-specific manner. All potentiation effects were found to last for periods of weeks. These findings demonstrate that both epileptogenic and non-epileptogenic piriform cortex stimulation induces lasting potentiation of population field responses in the entorhinal cortex of the awake rat. The LTP effects were inducible in a graded manner and were sensitive to the temporal context of stimulation. The finding that low-frequency activation of the septum can enhance plasticity in the entorhinal cortex adds to a body of data indicating a role for the medial septum in contributing to theta activity and plasticity in both the entorhinal cortex and hippocampal formation. Hippocampus 7:257–270, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

20.
Long-lasting modulation of intrinsic neuronal excitability in cortical neurons underlies distinct stages of skill learning. However, whether individual differences in learning capabilities are dependent on the rate by which such learning-induced modifications occur has yet to be explored. Here we show that training rats in a simple olfactory-discrimination task results in the same enhanced excitability in piriform cortex neurons as previously shown after training in a much more complex olfactory-discrimination task. Based on their learning capabilities in the simple task, rats could be divided to two groups: fast performers and slow performers. The rate at which rats accomplished the skill to perform the simple task was correlated with the time course at which piriform cortex neurons increased their repetitive spike firing. Twelve hours after learning, neurons from fast performers had reduced spike frequency adaptation as compared with neurons from slow performers and controls. Three days after learning, spike frequency adaptation was reduced in neurons from SP, while neurons from fast performers increased their spike firing adaptation to the level of controls. Accordingly, the post-burst AHP was reduced in neurons from fast performers 12 h after learning and in neurons from slow performers 3 days after learning. Moreover, the differences in learning capabilities between fast performers and slow performers were maintained when examined in a different, complex olfactory-discrimination task. We suggest that the rate at which neuronal excitability is modified during learning may affect the behavioral flexibility of the animal.  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