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.
This study describes the development of the ferret thalamic reticular nucleus from Nissl-stained and from parvalbumin-immunostained sections. From early stages [embryonic day (E) 23-E25], there is a large group of ventral thalamic cells which lies between the dorsal thalamus and the primordial internal capsule. This group of cells, the primordial reticular nucleus, gives rise to the main body of the reticular nucleus, the perigeniculate nucleus and the perireticular nucleus. In the reticular nucleus, there are two waves of parvalbumin expression during development. The first wave begins prenatally in small cells which are seen rarely after birth. Their fate is not clear: they may have lost immunoreactivity, migrated elsewhere, or died. At the end of the first wave, a second wave begins in a distinct group of larger ovoid reticular cells, which appear to remain into adulthood. At about birth, the dorsocaudal pole of the reticular nucleus first forms the perigeniculate nucleus. During this developmental stage, cells which make up the reticular and perigeniculate nuclei are the only parvalbumin-immunostained structures in the thalamus. Thus, rather than develop from the dorsal thalamus, the perigeniculate nucleus seems to have its origins in the ventral thalamus together with the reticular nucleus. During development, the reticular nucleus is associated closely with a large mass of cells located in the internal capsule, called the perireticular nucleus. Later, the perireticular nucleus is dramatically reduced in size: that is, there is a large reduction in the number of perireticular cells seen per section and in the extent of the nucleus across the internal capsule. There are two cytoarchitectonically distinct groups of perireticular cells. One group of cells, called the large-celled perireticular zone (LPR), enters the internal capsule from early prenatal development (E25). Many of these cells reach the globus pallidus and extend as far as the cortical subplate zone. The LPR together with the subplate form an extensive neuronal network in the white matter during early development, which disappears later in development (about postnatal day 20). The second group of perireticular cells is made up of smaller cells and is called the small-celled perireticular zone (SPR). These small cells enter the internal capsule from the reticular nucleus just prior to birth. Many of the cells in the SPR remain in the adult.  相似文献   

2.
We have examined the responses of a population of 77 cells in the dorsal lateral geniculate nucleus (dLGN) of the anaesthetized, paralysed cat. Here the synthetic enzyme for the production of nitric oxide, nitric oxide synthase, is found only in the presynaptic terminals of the cholinergic input from the brainstem. In our hands, iontophoretic application of inhibitors of this enzyme resulted both in significant decreases in visual responses and decreased responses to exogenous application of NMDA, effects which were reversed by coapplication of the natural substrate for nitric oxide synthase, L-arginine, but not the biologically inactive isomer, D-arginine. Nitroprusside and S -nitroso- N -acetylpenicillamine (SNAP), nitric oxide donors, but not L-arginine, were able to increase markedly both spontaneous activity and the responsiveness to NMDA application. Furthermore, SNAP application facilitated visual responses. Responses of cells in animals without retinal, cortical and parabrachial input to the LGN suggest a postsynaptic site of action of nitric oxide. This modulation of the gain of visual signals transmitted to the cortex suggests a completely novel pathway for nitric oxide regulation of function, as yet described only in primary sensory thalamus of the mammalian central nervous system.  相似文献   

3.
We have tested the effect of iontophoretic application of the nitric oxide synthase inhibitor l -nitroarginine on the activity of a population of 53 perigeniculate (PGN) cells, recorded extracellularly in the anaesthetized paralysed cat. In all cells tested with visual stimulation during l -nitroarginine application (n= 15), the visually elicited responses were markedly reduced, on average by 63 ± 15%, and there was a reduction in spontaneous activity too. This effect was blocked by co-application of the substrate for nitric oxide synthase, l -arginine, but not by the inactive d -isoform, although application of l -arginine alone was without effect. Pressure application of the nitric oxide donor S-nitroso-N-acetylpenicillamine (SNAP) elevated both visual responses and spontaneous discharge, an effect also seen with a second nitric oxide donor, sodium nitroprusside (n= 12). The nitric oxide synthase inhibitor l -nitroarginine was applied to a sub-population of seven cells and it selectively decreased NMDA mediated excitation (reduction 80 % 14%) with little or no effect on the excitation mediated by α-amino-3-hydroxy-5-5-methyl-4-isoxazole-propionic acid (AMPA) or quisqualate (effects not statistically significant), and it had no effect (n= 7) on excitation mediated by the metabotropic agonist (1S,3R)-1-aminocyclopentane-1,3-dicarboxylic acid (ACPD). Furthermore, application of SNAP also increased the magnitude of excitatory responses mediated by NMDA receptors. On a different population of seven cells, application of the new NO donor diethylamine-nitric oxide (DEA-NO) enhanced the actions Of NMDA without an effect on responses to AMPA. These effects are qualitatively and quantitatively similar to those we have previously described for X and Y type cells in the dorsal lateral geniculate nucleus (dLGN), despite the known opposite effects of acetylcholine (ACh) application in the dLGN and PGN (ACh is co-localized with nitric oxide synthase at both sites). We propose that within the PGN nitric oxide acts to enhance transmission utilizing NMDA receptors selectively (thereby interacting with the globally inhibiting effect of ACh at this site) to enhance visual responses, reducing or removing the non-specific inhibitory drive from PGN to dLGN seen in the spindling activity of slow-wave sleep. These effects will act in concert with the facilitatory actions of both ACh and nitric oxide within the dLGN proper, and will thereby enhance the faithful transmission of visual information from retina to cortex.  相似文献   

4.
The structural features of corticogeniculate axons were studied in adult cats after labeling them with horseradish peroxidase (HRP). Injections of HRP into the optic radiations near the dorsal lateral geniculate nucleus result in Golgi-like filling of both geniculate relay neurons and corticogeniculate axons. In the present material at least two main types of axons could be defined. The most common type is called the type I axon because it so closely resembles the type I axons described by Guillery ('66, '67) in Golgi preparations. These fine axons have smooth surfaces and consistent fiber diameter. Most terminal swellings are at the ends of short collateral branches and these swellings form asymmetric synaptic contacts onto small and medium-sized dendrites. Type I axons typically innervate more than one lamina as well as interlaminar zones and they clearly arise from the cerebral cortex. The second type of axon is called the beaded axon because of its numerous swellings, en passant. These swellings frequently are larger than those on type I axons and they differ from previously described corticogeniculate axon terminals in their ultrastructural features. That is, their synaptic contacts appear symmetrical and they form axosomatic contacts. Because of these differences, the possibility that beaded axons are of subcortical origin, particularly from the perigeniculate nucleus, is discussed. When type I axons and geniculate relay neurons are filled in the same region of the nucleus it is possible to identify probable sites of synaptic contact by using the light microscope. Such analyses indicate that corticogeniculate axons synapse directly onto relay cells, primarily on peripheral dendritic branches. Further, it appears that single axons contact many geniculate neurons and that single neurons are contacted by many axons.  相似文献   

5.
Terminal aborizations and synaptic boutons of cortical afferents to the cuneate nucleus were examined by light and electromicroscopy following intra-axonal staining with HRP. Two populations of afferents are described: (1) direct corticocuneate fibers, and (2) fibers destined for the spinal cord which issue collateral branches to the cuneate nucleus. Corticocuneate terminals primarily contact fine dendrites located in the ventral parts of the nucleus. These results are discussed in relation to previous anatomical findings and to new concepts of cuneate nucleus organization.  相似文献   

6.
Conventional light and confocal microscopy of thick vibratome sections of the hypothalamus of adult male and female rats immunostained for the astrocytic marker glial fibrillary acidic protein (GFAP) revealed that the supraoptic nucleus (SON) contains two morphologically distinct types of astrocytes. One has a stellate form, similar to that of most astrocytes in the adult CMS. The other has a morphology reminiscent of radial glia in the developing CNS: from their cell bodies, located along the ventral glia lamina (VGL), arise one long thick process that spans the SON in the coronal plane, several horizontally-oriented processes that form a dense network in the VGL, and a short process oriented towards the pia. The latter astrocytes are immunoreactive for vimentin, an intermediate filament protein of immature glial cells and a marker for radial glia. The stellate astrocytes showed no vimentin immunoreactivity. The functional significance of each type of supraoptic astrocyte is at present unknown but the presence of radial glia-like cells in this hypothalamic region suggests that the SON retains a certain degree of immaturity during adulthood, that may be linked to its well known capacity to undergo neuronal-glial plasticity under physiological and experimental stimulation.  相似文献   

7.
We investigated the source of axons and terminals in the cat's lateral geniculate nucleus that stain positively for NADPH-diaphorase. The functional significance of such staining is that NADPH-diaphorase is identical to the enzyme nitric oxide synthetase, and thus it is though to reveal cells and axons that use nitric oxide as a neuromodulator. Within the lateral geniculate and adjacent perigeniculate nuclei, a dense network of axons and terminals is labeled for NADPH-diaphorase, The pattern of NADPH-diaphorase staining here is remarkably similar to that of choline acetyltransferase (ChAT) staining, suggesting that the source of these axons and terminals might be the parabrachial region of the brainstem because this provides the major cholinergic input to the lateral geniculate nucleus. In other areas of the brain to which parabrachial axons project, there is also a similar staining pattern for NADPH-diaphorase and ChAT. Furthermore, the patterns of cell staining within the parabracial region for NADPH-diaphorase and ChAT are virtually identical. However, the relationship between ChAT and NADPH-diaphorase staining for the parabrachial region is not a general property of cholinergic neurons. Other cholinergic cells and axons, such as the trochlear nerve, the oculomotor nerve and nucleus, and the parabigeminal nucleus, which all label densely for ChAT, stain poorly or not at all for NADPH-diaphorase. It is significant that the parabigeminal nucleus, which provides a cholinergic input to the lateral geniculate nucleus, has no cells that label for NADPH-diaphorase. We used double labeling methods to identify further the source of NADPH-diaphorase staining in the lateral geniculate nucleus. We found that parabrachial cells co-localize NADPH-diaphorase and ChAT. Noradrenergic and serotoninergic cells in the brainstem also innervate the lateral geniculate nucleus, but we found that none of these co-localize NADPH-diaphorase. Finally, by combining NADPH-diaphorase histochemistry with retrograde labeling of cells that project to the lateral geniculate nucleus, we found that the cholinergic cells of the parabrachial region are essentially the sole source of NADPH-diaphorase in the lateral geniculate nucleus. We thus conclude that cells from the parabrachial region that innervate the lateral geniculate nucleus use both acetylcholine and nitric oxide for neurotransmission, and that this is virtually the only afferent input to this region that uses nitric oxide. © 1993 Wiley-Liss, Inc.  相似文献   

8.
Experiments were conducted with halothane-anesthetized cats implanted with a push-pull cannula in the caudate nucleus in order to estimate the effects of glutamate (GLU) agonists on the release of 3H-dopamine continuously synthesized from 3H-tyrosine. In the presence of tetrodotoxin (TTX), glutamate (10-8 M, 10-4 M) and kainate (KAI) (10-5 M) stimulated the release of 3H-dopamine while quisqualate (10-5 M) and N-methyl-D-aspartate (NMDA) (10-5 M) were without effect. The stimulatory effect of kainate (10-5 M) on 3H-dopamine release did not seem to be mediated by glutamate released from corticostriatal fibers, as not only kainate, but also quisqualate (QUI) and N-methyl-D-aspartate enhanced the efflux of glutamate through a tetrodotoxin-resistant process. Riluzole (10-5 M), gamma-D-glutamyl-glycine (GDGG) (10-5 M) and glutamine-diethyl-ester (10-5 M) prevented the stimulatory effect of kainate (10-5 M) while 6-cyano-7-nitro-quinoxaline-2,3-dione (CNQX) (10-5 M), kynurenate (10-5 M) and 2-amino-5-phosphonovalerate (APV) (10-5 M) were without effect. In the presence of concanavalin A (CONA) (10-7 M), a lectin which is known to prevent the quisqualate-evoked desensitization of glutamate receptors, quisqualate (10-5 M) stimulated the release of 3H-dopamine. In addition, in the absence of concanavalin A, quisqualate (10-5 M) blocked the stimulatory effects of kainate (10-5 M) or glutamate (10-4 M) on 3H-dopamine release. These results suggest the involvement of receptors of the quisqualate/kainate subtype in the direct glutamate-induced presynaptic facilitation of dopamine release. In contrast to what was observed in the presence of tetrodotoxin, in the absence of the neurotoxin, high concentrations of glutamate (10-4 M) and kainate (10-5 M) reduced rather than stimulated the release of 3H-dopamine. A weak inhibitory effect was also observed with quisqualate (10-5 M) while N-methyl-D-aspartate (10-5 M) was without effect. In the light of previous studies, these latter observations suggest that glutamate can also exert an indirect inhibitory presynaptic influence on the release of dopamine from nerve terminals of the nigrostriatal dopaminergic neurons by acting on receptors of the quisqualate/kainate subtype located on striatal GABAergic neurons.  相似文献   

9.
The dendrites of neocortical pyramidal cells are bombarded by myriads of synaptic inputs and express active conductances generating prominent plateau potentials. We have examined in vivo the possibility that spontaneous synaptic inputs trigger or terminate plateau potentials after blockage of K+ currents. Under barbiturate anaesthesia, pairs of cortical cells were intracellularly recorded with sharp electrodes from the cat's association cortex (areas 5–7). In one pyramidal cell, K+ channels were blocked with intracellular Cs+, while in the simultaneously impaled pyramidal cell the K+ conductances were left intact to act as a control; this second cell allowed recognition of spontaneous spindle-related synaptic activity. Depolarizing current pulses elicited single, all-or-none plateau potentials (60–70 mV, 0.1–0.4 s). Plateau potentials slowly repolarized towards a break point of fast repolarization around -20 mV. Thalamic-evoked inhibitory postsynaptic potentials consistently shut off the plateaus. Synchronized spontaneous activity, as occurring during thalamic-generated spindle oscillations, either triggered or blocked the plateaus. These results suggest that spontaneously occurring synaptic activation during synchronized oscillatory states, such as those that occur during sleep spindles in vivu , may exert a strong control over the dendritic excitability in neocortical pyramidal cells.  相似文献   

10.
The aim of this study was to extend the anatomical study of the corticoreticular organization in the monkey by means of microelectrophysiological techniques. Considering the relatively modest projection (see companion paper, Wiesendanger and Wiesendanger, '87), it was surprising to see that over 70% of the investigated LRN neurons were influenced from at least one cortical stimulation site. Many neurons responded, however, with long latencies suggesting an indirect transmission line. In line with the anatomical tracing study, most short-latency responses were obtained from the motor cortex. Postcentral cortex and the SMA were, in general, less effective sites for evoking responses in the LRN. LRN neurons with similar cortical inputs tended to be clustered together suggesting that the corticoreticular projection is discretely organized with an "intermingled somatotopy". The majority of the 87 tested LRN neurons were not reactive to any peripheral stimulus (33%) or responded only to nociceptive peripheral stimulation (31%). Very large receptive fields were seen in 8% of the units. However, a significant proportion of LRN neurons (10%) had restricted receptive fields and reacted to gentle cutaneous stimuli, and others (17%) responded to discrete passive rotations of one or more joints. There was often a somatotopical correspondence between the peripheral and the cortical inputs. It is concluded that the LRN in monkeys is under the influence of the motor cortex, which, however, may be exerted to a major extent via indirect pathways. The electrophysiological data suggest a discrete rather than a diffuse relationship with the LRN.  相似文献   

11.
A major inhibitory input to the dorsal thalamus arises from neurons in the thalamic reticular nucleus (TRN), which use gamma‐aminobutyric acid (GABA) as a neurotransmitter. We examined the synaptic targets of TRN terminals in the visual thalamus, including the A lamina of the dorsal lateral geniculate nucleus (LGN), the medial interlaminar nucleus (MIN), the lateral posterior nucleus (LP), and the pulvinar nucleus (PUL). To identify TRN terminals, we injected biocytin into the visual sector of the TRN to label terminals by anterograde transport. We then used postembedding immunocytochemical staining for GABA to distinguish TRN terminals as biocytin‐labeled GABA‐positive terminals and to distinguish the postsynaptic targets of TRN terminals as GABA‐negative thalamocortical cells or GABA‐positive interneurons. We found that, in all nuclei, the TRN provides GABAergic input primarily to thalamocortical relay cells (93–100%). Most of this input seems targeted to peripheral dendrites outside of glomeruli. The TRN does not appear to be a significant source of GABAergic input to interneurons in the visual thalamus. We also examined the synaptic targets of the overall population of GABAergic axon terminals (F1 profiles) within these same regions of the visual thalamus and found that the TRN contacts cannot account for all F1 profiles. In addition to F1 contacts on the dendrites of thalamocortical cells, which presumably include TRN terminals, another population of F1 profiles, most likely interneuron axons, provides input to GABAergic interneuron dendrites. Our results suggest that the TRN terminals are ideally situated to modulate thalamocortical transmission by controlling the response mode of thalamocortical cells. J. Comp. Neurol. 440:321–341, 2001. © 2001 Wiley‐Liss, Inc.  相似文献   

12.
In vitro intracellular recordings of membrane potential obtained from the oxytocin and vasopressin neurons of the mammalian hypothalamo-neurohypophysial system in slices (1–3) and expiants (4, 5) have demonstrated many of the intrinsic properties of these magnocellular neuroendocrine cells (MNCs). Voltage-clamp techniques, which are required to study directly the currents underlying intrinsic or transmitter-evoked potential changes, have been applied to cultured embryonic (6) or neonatal supraoptic neurons (7–9) and have been successfully applied to adult supraoptic neurons in situ in only one laboratory (10, 11). We have modified a technique for dissociation of viable adult guineapig hippocampal neurons (12) to dissociate supraoptic MNCs from adult rats for voltage-clamp studies. MNCs were selectively labelled with a fluorescent dye in vivo so that they could be identified after dissociation and prior to making recordings. These data have been published in abstract form elsewhere (13, 14).  相似文献   

13.
Small horseradish peroxidase injections in the perigeniculate nucleus (PGN) or the lateral geniculate nucleus (LGN) gave retrograde labeling of many cells in the pontomesencephalic reticular formation (RF), the nuclei raphe dorsalis and centralis linearis, locus coeruleus, nucleus of the optic tract and nucleus parabigeminalis. Antidromic stimulation was used to identify neurons in the RF projecting to the PGN-LGN complex. Threshold mapping through the PGN and the LGN shows separate projection from the reticular formation to the PGN and the LGN.  相似文献   

14.
15.
Accumulating evidence implicates the dorsomedial hypothalamic nucleus (DMH) in the regulation of autonomic and neuroendocrine stress responses. However, although projections from the DMH to the paraventricular hypothalamic nucleus (PVN), which is the critical site of the neuroendocrine stress axis, have been described, the impact of DMH neurones in the modulation of hypothalamic‐pituitary‐adrenal (HPA) axis activation during stress is not fully understood. The present study aimed to investigate the role of the DMH in HPA axis responses to different types of stimuli. Male Sprague–Dawley rats fitted with a chronic jugular venous catheter were exposed to either an emotional stressor (elevated platform‐exposure) or immune challenge (systemic interleukin‐1β administration). Bilateral electrolytic lesions of the DMH disinhibited HPA axis responses to the emotional stressor, as indicated by higher plasma adrenocorticotrophic hormone levels during and after elevated platform exposure in lesioned animals compared to sham‐lesioned controls. Moreover, DMH‐lesioned animals showed increased neuronal activation in the PVN, as indicated by a higher c‐Fos expression after elevated‐platform exposure compared to controls. By contrast, DMH‐lesions had no effects on HPA axis responses to immune challenge. Taken together, our data suggest an inhibitory role of DMH neurones on stress‐induced HPA axis activation that is dependent upon the nature of the stimulus being important in response to an emotional stressor but not to immune challenge.  相似文献   

16.
D Demêmes  J Raymond 《Brain research》1982,231(2):433-437
Slices of the cat third oculomotor nucleus were incubated in vitro with [3H]glutamic acid. Electron microscopic radioautographs revealed that glutamate had been taken by small nerve endings distributed on the oculomotor motoneuron distal dendrites. In contrast, there was no uptake in the other types of terminals. The labeled terminals seem to correspond to the excitatory vestibulo-oculomotor nerve endings and different correlations suggest their glutamergic nature.  相似文献   

17.
18.
The bed nucleus of the stria terminalis (BNST) is known to play a critical role in mediating the behavioural and autonomic responses to stressors. The oval nucleus of the BNST (BNSTov) contains cell bodies that synthesise the stress hormone corticotropin‐releasing factor (CRF). Although afferent fibres originating from the BNSTov have been shown to innervate several key structures of the neuroendocrine and central autonomic system, the question remains as to whether some of these fibres are CRF‐positive. To directly address this question, we injected a ‘floxed’ anterograde tracer (rAAV5/EF1a‐DIO‐mCherry) into the BNSTov of CRFp3.0CreGFP transgenic mice, which express a green fluorescent protein (GFP) under the control of the CRF promoter. Serial sections were then analysed for the presence of double‐labelled fibres in potential projection sites. To determine whether CRF neurons in the rat BNSTov send comparable projections, we infused rat BNSTov with an adeno‐associated viral vector (AAV) in which the human synapsin promoter drives enhanced GFP expression. We then used CRF immunoreactivity to examine double‐labelled fluorescent fibres and axon terminals in projection sites from brain sections of the AAV‐infused rats. We have observed several terminal fields in the mouse and rat brain with double‐labelled fibres in the Dorsal raphe nucleus (DRD), the paraventricular nucleus of the hypothalamus and, to a lesser extent, in the ventral tegmental area. We found double‐labelled terminal boutons in the nucleus accumbens shell, prelimbic cortex and posterior basolateral nucleus of the amygdala. The most intense double‐labelling was found in midbrain, including substantia nigra pars compacta, red nucleus, periaqueductal grey and pontine nuclei, as well as DRD. The results of the present study indicate that CRF neurons are the output neurons of the BNSTov and they send projections not only to the centres of neuroendocrine and autonomic regulation, but also regions modulating reward and motivation, vigilance and motor function, as well as affective behaviour.  相似文献   

19.
Quantitative in situ hybridization histochemistry was used to examine the effects of unilateral 6–hydroxydopamine lesions of the ascending dopaminergic fibres on levels of mRNA encoding the neuropeptides enkephalin, dynorphin and substance P in subregions of the nucleus accumbens. The nucleus accumbens was divided into quadrants and changes in mRNA were measured along the rostrocaudal extent of the nucleus. Two weeks after the lesion an increase was found in enkephalin mRNA in the lesioned side compared to the non-lesioned side, whereas a decrease was observed for dynorphin and substance P mRNA. The changes in mRNA levels differed from quadrant to quadrant and were not uniformly distributed along the rostrocaudal axis. Both types of changes, i.e. increase and decrease, were much higher in rostral parts of the nucleus than in caudal parts, indicating regional differences in the effects of blockade of the dopaminergic neurotransmission. The lesion-induced increases and decreases in mRNA levels occurred in both the shell and the core subregions of the nucleus accumbens and were not specifically related to either of these areas. Factors are discussed that may contribute to the rostrocaudal gradient in the changes of enkephalin, substance P and dynorphin mRNA levels. On the basis of their afferent and efferent connections, the rostral and caudal parts of the nucleus accumbens are considered to be involved in different functions. The present results suggest that dopamine depletion may affect these functions in a differential manner.  相似文献   

20.
Projections to and from the visual sector of the thalamic reticular nucleus were studied in the prosimian primate genus Galago by anterograde and retrograde transport of WGA-HRP injected into the dorsal lateral geniculate nucleus (GLd), pulvinar nucleus, and their cortical targets. Contrary to the idea that thalamic connections with the reticular nucleus are not delimited sharply between nuclei associated with the same modality, our results show a distinct laminar segregation of the projections from the GLd and pulvinar nuclei. The GLd is connected reciprocally with the lateral {frsol|2/3} of the caudal part of the reticular nucleus, and the striate cortex sends projections to the same lateral tier. Both sets of projections are organized topographically, lines of projection taking the form of slender elongated strips that run from caudo-dorsal to rostro-ventral within the nucleus. The pulvinar nucleus, which projects to several areas of the temporal, parietal, and occipital lobes, including the striate cortex, is connected reciprocally with the medial {frsol|1/3} of the caudal part of the reticular nucleus. Every injection into the pulvinar nucleus labelled a wide area of the medial tier, with no indication of visuotopic organization. The projections from the middle temporal area, one of the principal targets of the pulvinar nucleus, also terminate only in the medial tier of the visual sector. And we would expect that, in general, a thalamic nucleus and its cortical target would project to the same part of the reticular nucleus. The case of the striate area is an exception but only in the sense that it projects to the pulvinar nucleus as well as GLd. Thus an injection into a single locus in area 17 produces two parallel strips in the visual sector of the reticular nucleus, but both are in the lateral tier. We propose that each strip arises from a separate population of cells with cortical layer VI, one with an allegiance to the GLd and the other to the pulvinar nucleus.  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