Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-p2v8j Total loading time: 0 Render date: 2024-05-04T15:14:29.878Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  18 August 2009

Mircea Steriade
Affiliation:
Université Laval, Québec
Denis Pare
Affiliation:
Rutgers University, New Jersey
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2007

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Abel, T., Nguyen, P. V., Barad, M.et al. (1997) Genetic demonstration of a role for PKA in the late phase of LTP and in hippocampus-based long-term memory. Cell 88: 615–26.CrossRefGoogle ScholarPubMed
Abeles, M., Bergman, H., Gat, I.et al. (1995) Cortical activity flips among quasi-stationary states. Proceedings of the National Academy of Sciences USA 92: 8616–20.CrossRefGoogle ScholarPubMed
Achermann, P. and Borbély, A. (1997) Low-frequency (<1 Hz) oscillations in the human sleep EEG. Neuroscience 81: 213–22.CrossRefGoogle Scholar
Adams, J. H., Graham, D. I. and Jennett, B. (2000) The neuropathology of the vegetative state after an acute brain insult. Brain 123: 1327–38.CrossRefGoogle ScholarPubMed
Adolphs, R., Cahill, L., Schul, R. and Babinsky, R. (1997) Impaired declarative memory for emotional stimuli following bilateral amygdala damage in humans. Learning and Memory 4: 291–300.CrossRefGoogle ScholarPubMed
Aggleton, J. P. (2000) The Amygdala: A Functional Analysis. Oxford: Oxford University Press.Google Scholar
Aggleton, J. P. and Mishkin, M. (1983a) Visual recognition impairment following medial thalamic lesions. Neuropsychologia 21: 189–97.CrossRefGoogle Scholar
Aggleton, J. P. and Mishkin, M. (1983b) Memory impairments following restricted medial thalamic lesions in monkeys. Experimental Brain Research 52: 199–209.CrossRefGoogle Scholar
Aggleton, J. P., Hunt, P. R. and Rawlins, J. N. (1986) The effects of hippocampal lesions upon spatial and non-spatial tests of working memory. Behavioral Brain Research 19: 133–46.CrossRefGoogle ScholarPubMed
Aghajanian, G. K. (1985) Modulation of a transient outward current in serotonergic neurones by alpha 1-adrenoceptors. Nature 315: 501–3.CrossRefGoogle ScholarPubMed
Aghajanian, G. K. and Wang, E. Y. (1977) Habenular and other midbrain raphe afferents demonstrated by a modified retrograde tracing technique. Brain Research 122: 229–42.CrossRefGoogle ScholarPubMed
Agranoff, B. W., Davis, R. E. and Brink, J. J. (1966) Chemical studies on memory fixation in goldfish. Brain Research 1: 303–9.CrossRefGoogle ScholarPubMed
Airaksinen, M. S. and Panula, P. (1988) The histaminergic system in the guinea pig central nervous system: an immunocytochemical mapping study using an antiserum against histamine. Journal of Comparative Neurology 273: 163–86.CrossRefGoogle ScholarPubMed
Aitkin, L. M., Irvine, D. R., Nelson, J. E., Merzenich, M. M. and Clarey, J. C. (1986) Frequency representation in the auditory midbrain and forebrain of a marsupial, the northern native cat(Dasyurus hallucatus). Brain Behavior and Evolution 29: 17–28.CrossRefGoogle Scholar
Alden, M., Besson, J. M. and Bernard, J. F. (1994) Organization of the efferent projections from the pontine parabrachial area to the bed nucleus of the stria terminalis and neighboring regions: a PHA-L study in the rat. Journal of Comparative Neurology 341: 289–314.CrossRefGoogle ScholarPubMed
Aleksanov, S. N. (1983) Coherent functions of the electrical activity of the hippocampus, amygdala and frontal cortex during alimentary instrumental reflexes in the dog. Zhurnal Vysshei Nervnoi Deyatelnosti Imeni I P Pavlova 33: 694–9.Google ScholarPubMed
Allison, T., McCarthy, G., Wood, C. C., Williamson, P. D. and Spencer, D. D. (1989) Human cortical potentials evoked by stimulation of the median nerve. II. Cytoarchitectonic areas generating long-latency activity. Journal of Neurophysiology 62: 711–22.CrossRefGoogle ScholarPubMed
Alloway, K. D., Wallace, M. B. and Johnson, M. J. (1994) Cross-correlation analysis of cuneothalamic interactions in the rat somatosensory system: influence of receptive field topography and comparisons with thalamocortical interactions. Journal of Neurophysiology 72: 1949–72.CrossRefGoogle ScholarPubMed
Alonso, A. (2002) Electrophysiology of neurones in the perirhinal and entorhinal cortices and neuromodulatory changes in firing patterns. In The Parahippocampal Region, ed. Witter, M. P. and Wouterlood, F., pp. 89–105. Oxford: Oxford University Press.Google Scholar
Alonso, A. and Garcia-Austt, E. (1987a) Neuronal sources of theta rhythm in the enthorinal cortex of the rat. I. Laminar distribution of theta field potentials. Experimental Brain Research 67: 493–501.CrossRefGoogle Scholar
Alonso, A. and Garcia-Austt, E. (1987b) Neuronal sources of theta rhythm in the enthorinal cortex of the rat. II. Phase relations between unit discharges and theta field potentials. Experimental Brain Research 67: 502–9.CrossRefGoogle Scholar
Alonso, A. and Klink, R. (1993) Differential electroresponsiveness of stellate and pyramidal-like cells of medial entorhinal cortex layer II. Journal of Neurophysiology 70: 128–43.CrossRefGoogle ScholarPubMed
Alonso, A. and Köhler, C. (1984) A study of the reciprocal connections between the septum and the entorhinal area using anterograde and retrograde axonal transport methods in the rat brain. Journal of Comparative Neurology 225: 327–43.CrossRefGoogle ScholarPubMed
Alonso, A. and Llinás, R. (1989) Subthreshold theta-like rhythmicity in stellate cells of enthorinal cortex layer II. Nature 342: 175–7.CrossRefGoogle Scholar
Alonso, A., Faure, M. F. and Beaudet, A. (1994) Neurotensin promotes oscillatory bursting behaviour and is internalized in basal forebrain cholinergic neurons. Journal of Neuroscience 14: 5778–92.CrossRefGoogle ScholarPubMed
Alonso, A., Khateb, A., Fort, P., Jones, B. E. and Mühlethaler, M. (1996) Differential oscillatory properties of cholinergic and noncholinergic nucleus basalis neurons in guinea pig brain slices. European Journal of Neuroscience 8: 169–82.CrossRefGoogle Scholar
Amaral, D. G. (1986) Amygdalohippocampal and amygdalocortical projections in the primate brain. Advances in Experimental Medicine and Biology 203: 3–17.CrossRefGoogle ScholarPubMed
Amaral, D. G. and Cowan, W. M. (1980) Subcortical afferents to the hippocampal formation in the monkey. Journal of Comparative Neurology 189: 573–91.CrossRefGoogle ScholarPubMed
Amaral, D. G. and Price, J. L. (1984) Amygdalo-cortical projections in the monkey(Macaca fascicularis). Journal of Comparative Neurology 230: 465–96.CrossRefGoogle Scholar
Amaral, D. G., Price, J. L., Pitkänen, A. and Carmichael, S. T. (1992) Anatomical organization of the primate amygdaloid complex. In The Amygdala: Neurobiological Aspects of Emotion, Memory, and Mental Dysfunction, ed. Aggleton, J. P., pp. 1–66. New York: Wiley-Liss.Google Scholar
Amorapanth, P., LeDoux, J. E. and Nader, K. (2000) Different lateral amygdala outputs mediate reactions and actions elicited by a fear-arousing stimulus. Nature Neuroscience 3: 74–9.CrossRefGoogle ScholarPubMed
Amzica, F. and Massimini, M. (2002) Glial and neuronal interactions during slow waves and paroxysmal activities in the neocortex. Cerebral Cortex 12: 1101–13.CrossRefGoogle Scholar
Amzica, F. and Neckelmann, D. (1999) Membrane capacitance of cortical neurons and glia during sleep oscillations and spike-wave seizures. Journal of Neurophysiology 82: 2731–46.CrossRefGoogle ScholarPubMed
Amzica, F. and Steriade, M. (1995a) Disconnection of intracortical synaptic linkages disrupts synchronization of a slow oscillation. Journal of Neuroscience 15: 4658–77.CrossRefGoogle Scholar
Amzica, F. and Steriade, M. (1995b) Short- and long-range neuronal synchronization of the slow (<1 Hz) cortical oscillation. Journal of Neurophysiology 75: 20–38.CrossRefGoogle Scholar
Amzica, F. and Steriade, M. (1996) Progressive cortical synchronization of ponto-geniculo-occipital potentials during rapid eye movement sleep. Neuroscience 72: 309–14.CrossRefGoogle ScholarPubMed
Amzica, F. and Steriade, M. (1997) The K-complex: its slow (<1 Hz) rhythmicity and relation to delta waves. Neurology 49: 952–9.CrossRefGoogle ScholarPubMed
Amzica, F. and Steriade, M. (1998a) Cellular substrates and laminar profile of sleep K-complex. Neuroscience 82: 671–86.CrossRefGoogle Scholar
Amzica, F. and Steriade, M. (1998b) Electrophysiological correlates of sleep delta waves. Electroencephalography and Clinical Neurophysiology 107: 69–83.CrossRefGoogle Scholar
Amzica, F. and Steriade, M. (2000) Neuronal and glial membrane potentials during sleep and paroxysmal oscillations in the cortex. Journal of Neuroscience 20: 6646–65.CrossRefGoogle Scholar
Amzica, F. and Steriade, M. (2002) The functional significance of K-complexes. Sleep Medicine Reviews 6: 139–49.CrossRefGoogle ScholarPubMed
Amzica, F., Neckelmann, D. and Steriade, M. (1997) Instrumental conditioning of fast (20- to 50 Hz) oscillations in corticothalamic networks. Proceedings of the National Academy of Sciences USA 94: 1985–9.CrossRefGoogle ScholarPubMed
Amzica, F., Massimini, M. and Manfridi, A. (2002) Spatial buffering during slow and paroxysmal oscillations in cortical networks of glial cellsin vivo. Journal of Neuroscience 22: 1042–53.CrossRefGoogle ScholarPubMed
Andersen, P. and Andersson, S. A. (1968) Physiological Basis of Alpha Rhythm. New York: Appleton-Century-Crofts.Google Scholar
Anderson, A. K. and Phelps, E. A. (2001) Lesions of the human amygdala impair enhanced perception of emotionally salient events. Nature 411: 305–9.CrossRefGoogle ScholarPubMed
Andreasen, N. C., Arndt, S., Swayze, V.et al. (1994) Thalamic abnormalities in schizophrenia visualied through magnetic resonance image averaging. Science 266: 294–8.CrossRefGoogle Scholar
Arieli, A., Shoham, D., Hildesheim, R. and Grinvald, A. (1995) Coherent spatiotemporal patterns of ongoing activity revealed by real-time optical imaging coupled with single-unit recording in the cat visual cortex. Journal of Neurophysiology 73: 2072–93.CrossRefGoogle ScholarPubMed
Arieli, A., Sterkin, A., Grinvald, A. and Aertsen, A. (1996) Dynamics of ongoing activity: explanation of the large variability in evoked cortical responses. Science 273: 1868–71.CrossRefGoogle ScholarPubMed
Asanuma, C. (1992) Noradrenergic innervation of the thalamic reticular nucleus: a light and electron microscopic immunohistochemical study in rats. Journal of Comparative Neurology 319: 299–311.CrossRefGoogle ScholarPubMed
Asanuma, C. (1997) Distribution of neuromodulatory inputs in the reticular and dorsal thalamic nuclei. In Thalamus, vol. 2 (Experimental and Clinical Aspects), ed. Steriade, M., Jones, E. G. and McCormick, D. A., pp. 93–153. Oxford: Elsevier.Google Scholar
Asanuma, C. and Porter, L. L. (1990) Light and electron microscopic evidence for a GABAergic projection from the caudal basal forebrian to the thalamic reticular nucleus in rats. Journal of Comparative Neurology 302: 159–72.CrossRefGoogle ScholarPubMed
Aston-Jones, G., Ennis, M., Pieribone, V. A., Nickell, W. T. and Shipley, M. T. (1986) The brain nucleus locus coeruleus: Restricted afferent control of a broad efferent network. Science 234: 734–7.CrossRefGoogle ScholarPubMed
Bacci, A., Huguenard, J. R. and Prince, D. A. (2003) Functional autaptic neurotransmission in fast-spiking interneurons: a novel form of feedback inhibition in the neocortex. Journal of Neuroscience 23: 859–66.CrossRefGoogle ScholarPubMed
Bahar, A., Samuel, A., Hazvi, S. and Dudai, Y. (2003) The amygdalar circuit that acquires taste aversion memory differs from the circuit that extinguishes it. European Journal of Neuroscience 17: 1527–30.CrossRefGoogle Scholar
Bailey, D. J., Kim, J. J., Sun, W., Thompson, R. F. and Helmstetter, F. J. (1999) Acquisition of fear conditioning in rats requires the synthesis of mRNA in the amygdala. Behavioral Neuroscience 113: 276–82.CrossRefGoogle ScholarPubMed
Baker, S. N., Olivier, E. and Lemon, R. N. (1997) Coherent oscillations in the monkey motor cortex and hand muscle EMG show task-dependent modulation. Journal of Physiology (London) 501: 225–41.CrossRefGoogle ScholarPubMed
Bal, T. and McCormick, D. A. (1996) What stops synchronized thalamocortical oscillations?Neuron 17: 297–308.CrossRefGoogle ScholarPubMed
Bal, T., Krosigk, M. and McCormick, D. A. (1995a) Synaptic and membrane mechanisms underlying synchronized oscillations in the ferret lateral geniculate nucleusin vitro. Journal of Physiology (London) 483: 641–63.CrossRefGoogle Scholar
Bal, T., Krosigk, M. and McCormick, D. A. (1995b) Role of the ferret perigeniculate nucleus in the generation of synchronized oscillationsin vitro. Journal of Physiology (London) 483: 665–85.CrossRefGoogle Scholar
Ball, G. J., Gloor, P. and Schaul, N. (1977) The cortical electromicrophysiology of pathological delta waves in the electroencephalogram of cats. Electroencephalography and Clinical Neurophysiology 43: 346–61.CrossRefGoogle ScholarPubMed
Barazangi, N. and Role, L. W. (2001) Nicotine-induced enhancement of glutamatergic and GABAergic synaptic transmission in the mouse amygdala. Journal of Neurophysiology 86: 463–74.CrossRefGoogle ScholarPubMed
Barbas, H. and Olmos, D, J. S. (1990) Projections from the amygdala to basoventral and mediodorsal prefrontal regions in the rhesus monkey. Journal of Comparative Neurology 300: 549–71.CrossRefGoogle ScholarPubMed
Bargas, J., Galarraga, E. and Aceves, J. (1989) An early outward conductance modulates the firing latency and frequency of neostriatal neurons of the rat brain. Experimental Brain Research 75: 146–56.CrossRefGoogle ScholarPubMed
Barria, A., Muller, D., Derkach, V., Griffith, L. C. and Soderling, T. R. (1997) Regulatory phosphorylation of AMPA-type glutamate receptors by CaM-KII during long-term potentiation. Science 276: 2042–5.CrossRefGoogle ScholarPubMed
Batsel, H. L. (1964) Spontaneous desynchronization in the chronic cat ‘cerveau isolé’. Archives Italiennes de Biologie 102: 547–66.Google Scholar
Battaglia, F. P., Sutherland, G. R. and McNaughton, B. L. (2004) Hippocampal sharp wave bursts coincide with neocortical ‘up-state’ transitions. Learning and Memory 11: 697–704.CrossRefGoogle ScholarPubMed
Bauer, E. P. and LeDoux, J. E. (2004) Heterosynaptic long-term potentiation of inhibitory interneurons in the lateral amygdala. Journal of Neuroscience 24: 9507–12.CrossRefGoogle ScholarPubMed
Bauer, E. P., LeDoux, J. E. and Nader, K. (2001) Fear conditioning and LTP in the lateral amygdala are sensitive to the same stimulus contingencies. Nature Neuroscience 4: 687–8.CrossRefGoogle ScholarPubMed
Baughman, R. W. and Gilbert, C. D. (1980) Aspartate and glutamate as possible neurotransmitters of cells in layer 6 of the visual cortex. Nature 287: 848–50.CrossRefGoogle ScholarPubMed
Baxter, M. G. and Murray, E. A. (2002) The amygdala and reward. Nature Reviews Neuroscience 3: 563–73.CrossRefGoogle ScholarPubMed
Bayer, L., Eggermann, E., Saint-Mieux, B.et al. (2002) Selective action of orexin (hypocretin) on nonspecific thalamocortical projection neurons. Journal of Neuroscience 22: 7835–9.CrossRefGoogle ScholarPubMed
Bayer, L., Serafin, M., Eggermann, E.et al. (2004) Exclusive postsynaptic action of hypocretin-orexin on sublayer 6b cortical neurons. Journal of Neuroscience 24: 6760–4.CrossRefGoogle ScholarPubMed
Bazhenov, M., Timofeev, I., Steriade, M. and Sejnowski, T. J. (1998a) Cellular and network models for intrathalamic augmenting responses during 10-Hz stimulation. Journal of Neurophysiology 79: 2730–48.CrossRefGoogle Scholar
Bazhenov, M., Timofeev, I., Steriade, M. and Sejnowski, T. J. (1998b) Computational models of thalamocortical augmenting responses. Journal of Neuroscience 18: 6444–65.CrossRefGoogle Scholar
Bazhenov, M., Timofeev, I., Steriade, M. and Sejnowski, T. J. (1999) Self-sustained rhythmic activity in the thalamic reticular nucleus mediated by depolarizing GABAA receptor potentials. Nature Neuroscience 2: 168–74.CrossRefGoogle ScholarPubMed
Bazhenov, M., Timofeev, I., Steriade, M. and Sejnowski, T. (2000) Spiking-bursting activity in the thalamic reticular nucleus initiates sequences of spindle oscillations in thalamic networks. Journal of Neurophysiology 84: 1076–87.CrossRefGoogle ScholarPubMed
Bazhenov, M., Timofeev, I., Steriade, M. and Sejnowski, T. J. (2002) Model of thalamocortical slow-wave sleep oscillations and transitions to activated states. Journal of Neuroscience 22: 8691–704.CrossRefGoogle ScholarPubMed
Bear, M. F. and Abraham, W. C. (1996) Long-term depression in hippocampus. Annual Review of Neuroscience 19: 437–62.CrossRefGoogle ScholarPubMed
Beaudet, A., and Descarries, L. (1978) The monoamine innervation of rat cerebral cortex: synaptic and non-synaptic axon terminals. Neuroscience 3: 851–60.CrossRefGoogle Scholar
Bechara, A., Tranel, D., Damasio, H.et al. (1995) Double dissociation of conditioning and declarative knowledge relative to the amygdala and hippocampus in humans. Science 269: 1115–18.CrossRefGoogle ScholarPubMed
Beckstead, R. M. (1978) Afferent connections of the entorhinal area in the rat as demonstrated by retrograde cell-labeling with horse radish peroxidase. Brain Research 152: 249–64.CrossRefGoogle Scholar
Beggs, J. M. and Kairiss, E. W. (1994) Electrophysiology and morphology of neurons in rat perirhinal cortex. Brain Research 665: 18–32.CrossRefGoogle ScholarPubMed
Behrendt, R. P. (2003) Hallucinations: synchronization of thalamocortical gamma oscillations underconstrained by sensory input. Consciousness and Cognition 12: 413–51.CrossRefGoogle Scholar
Bellgowan, P. S. and Helmstetter, F. J. (1996) Neural systems for the expression of hypoalgesia during nonassociative fear. Behavioral Neuroscience 110: 727–36.CrossRefGoogle ScholarPubMed
Ben-Ari, Y., Gal La Salle, G., Barbin, G., Schwartz, J. C. and Garbarg, M. (1977) Histamine synthesizing afferents within the amygdaloid complex and bed nucleus of the stria terminalis of the rat. Brain Research 138: 285–94.CrossRefGoogle ScholarPubMed
Bennett, M. V. L. (1970) Comparative physiology: electric organs. Annual Review of Physiology 32: 471–528.CrossRefGoogle ScholarPubMed
Bentivoglio, M., Aggleton, J. P. and Mishkin, M. (1997) The thalamus and memory formation. In Thalamus, vol. 2 (Experimental and Clinical Aspects), ed. Steriade, M., Jones, E. G. & McCormick, D. A., pp. 689–720. Oxford: Elsevier.
Berger, B., Thierry, A. M., Tassin, J. P. and Moyne, M. A. (1976) Dopaminergic innervation of the rat prefrontal cortex: a fluorescence histochemical study. Brain Research 106: 133–45.CrossRefGoogle ScholarPubMed
Berger, H. (1929) Uber das Elektroencephalogramm des Menschen. Archive für Psychiatrie und Nervenkrankheiten 87: 527–70.CrossRefGoogle Scholar
Berger, H. (1937) Uber das Elektroencephalogramm des Menschen. Dreizehnte Mitteilung. Archive für Psychiatrie und Nervenkrankheiten 106: 577–84.CrossRefGoogle Scholar
Berger, T., Senn, W. and Lüscher, H. R. (2003) Hyperpolarization-activated current Ih disconnects somatic and dendritic spike initiation zones in layer V pyramidal neurons. Journal of Neurophysiology 90: 2428–37.CrossRefGoogle ScholarPubMed
Berlucchi, G. (1997) One or many arousal systems? Reflections on some of Giuseppe Moruzzi's foresights and insights about intrinsic regulation of brain activity. Archives Italiennes de Biologie 135: 5–14.Google ScholarPubMed
Bernard, J. F., Huang, G. F. and Besson, J. M. (1990) Effect of noxious stimulation on the activity of neurons of the nucleus centralis of the amygdala. Brain Research 523: 347–50.CrossRefGoogle ScholarPubMed
Bernard, J. F., Huang, G. F. and Besson, J. M. (1992) Nucleus centralis of the amygdala and the globus pallidus ventralis: electrophysiological evidence for an involvement in pain processes. Journal of Neurophysiology 68: 551–69.CrossRefGoogle ScholarPubMed
Bernard, J. F., Alden, M. and Besson, J. M. (1993) The organization of the efferent projections from the pontine parabrachial area to the amygdaloid complex: A Phaseolus vulgaris leucoagglutinin (PHA-L) study in the rat. Journal of Comparative Neurology 329: 201–29.CrossRefGoogle ScholarPubMed
Berretta, S., Pantazopoulos, H., Caldera, M., Pantazopoulos, P. and Paré, D. (2005) Infralimbic cortex activation increases c-fos expression in intercalated neurons of the amygdala. Neuroscience 132: 943–53.CrossRefGoogle ScholarPubMed
Bickford, M. E., Günlük, A. E., Guido, W. and Sherman, S. M. (1993) Evidence that cholinergic axons from the parabrachial region of the brainstem are the exclusive source of nitric oxide in the lateral geniculate nucleus of the cat. Journal of Comparative Neurology 334: 410–30.CrossRefGoogle ScholarPubMed
Biella, G., Uva, L. and Curtis, M. (2001) Network activity evoked by neocortical stimulation in area 36 of the guinea pig perirhinal cortex. Journal of Neurophysiology 86: 164–72.CrossRefGoogle ScholarPubMed
Biella, G., Uva, L. and Curtis, M. (2002) Propagation of neuronal activity along the neocortical-perirhinal-entorhinal pathway in the guinea pig. Journal of Neuroscience 22: 9972–9.CrossRefGoogle ScholarPubMed
Bilkey, D. K. (1996) Long-term potentiation in the in vitro perirhinal cortex displays associative properties. Brain Research 733: 297–300.CrossRefGoogle ScholarPubMed
Bissière, S., Humeau, Y. and Lüthi, A. (2003) Dopamine gates LTP induction in lateral amygdala by suppressing feedforward inhibition. Nature Neuroscience 6: 587–92.CrossRefGoogle ScholarPubMed
Bland, B. H. (1986) The physiology and pharmacology of hippocampal formation theta rhythms. Progress in Neurobiology 26: 1–54.CrossRefGoogle ScholarPubMed
Bland, B. H., Andersen, P., Ganes, T. and Sven, O. (1980) Automated analysis of rhythmicity of physiologically identified hippocampal formation neurons. Experimental Brain Research 38: 205–19.CrossRefGoogle ScholarPubMed
Bliss, T. and Collingridge, G. L. (1993) A synaptic model of memory: long-term potentiation in the hippocampus. Nature 361: 31–9.CrossRefGoogle ScholarPubMed
Boejinga, P. H. and Lope, da Silva F. H. (1988) Differential distribution of beta and theta EEG activity in the entorhinal cortex of the cat. Brain Research 448: 272–86.CrossRefGoogle Scholar
Bordi, F., LeDoux, J., Clugnet, M. C. and Pavlides, C. (1993) Single-unit activity in the lateral nucleus of the amygdala and overlying areas of the striatum in freely behaving rats: rates, discharge patterns and responses to acoustic stimuli. Behavioral Neuroscience 107: 757–69.CrossRefGoogle ScholarPubMed
Bouyer, J. J., Montaron, M. F., Vahnée, J. M., Albert, M. P. and Rougeul, A. (1987) Anatomical localization of cortical beta rhythm in cat. Neuroscience 22: 863–9.CrossRefGoogle ScholarPubMed
Braga, M. F., Aroniadou-Anderjaska, V., Manion, S. T., Hough, C. J. and Li, H. (2004) Stress impairs alpha(1A) adrenoceptor-mediated noradrenergic facilitation of GABAergic transmission in the basolateral amygdala. Neuropsychopharmacology 29: 45–58.CrossRefGoogle ScholarPubMed
Bragin, A., Jandó, G., Nádasdy, Z.et al. (1995a) Gamma (40–100 Hz) oscillation in the hippocampus of the behaving rat. Journal of Neuroscience 15: 47–60.CrossRefGoogle Scholar
Bragin, A., Jandó, G., Nádasdy, Z., Van, L. M. and Buzsáki, G. (1995b) Dentate EEG spikes and associated interneuronal population bursts in the hippocampal hilar region of the rat. Journal of Neurophysiology 73: 1691–705.CrossRefGoogle Scholar
Braitenberg, V. and Schütz, A. (1991) Anatomy of the Cortex. Berlin: Springer.CrossRefGoogle Scholar
Braun, A. R., Balkin, T. J., Wesensten, N. J.et al. (1997) Regional cerebral blood flow throughout the sleep-wake cycle. Brain 120: 1173–97.CrossRefGoogle ScholarPubMed
Brazier, M. A. B. (1961) A History of the Electrical Activity of the Brain. London: Pitman.Google Scholar
Brazier, M. A. B. (1968) Studies of the EEG activity of limbic structures in man. Electroencephalography and Clinical Neurophysiology 25: 309–18.CrossRefGoogle ScholarPubMed
Bremer, F. (1935) Cerveau ‘isolé’ et physiologie du sommeil. Comptes Rendus de la Société de Biologie (Paris) 118: 1235–41.Google Scholar
Bremer, F. and Stoupel, N. (1959) Facilitation et inhibition des potentiels évoqués corticaux dans l'éveil cérébral. Archives Internationales de Physiologie 67: 240–75.CrossRefGoogle Scholar
Bremer, F. and Terzuolo, C. (1954) Contribution à l'étude des mécanismes physiologiques du maintien de l'activité vigile du cerveau. Interaction de la formation réticulée et de l'écorce cérébrale dans le processus de l'éveil. Archives Internationales de Physiologie 62: 157–78.Google Scholar
Bremer, F., Stoupel, N. and Reeth, P. C. (1960) Nouvelles recherches sur la facilitation et l'inhibition des potentiels évoqués corticaux dans l'éveil réticulaire. Archives Italiennes de Biologie 98: 229–47.Google Scholar
Bremner, J. D., Staib, L. H., Kaloupek, D.et al. (1999) Neural correlates of exposure to traumatic pictures and sound in Vietnam combat veterans with and without posttraumatic stress disorder: A positron emission tomography study. Biological Psychiatry 45: 806–16.CrossRefGoogle ScholarPubMed
Bringuier, V., Frégnac, Y., Baranyi, A., Debanne, D. and Shulz, D. E. (1997) Synaptic origin and stimulus dependency of neuronal oscillatory activity in the primary visual cortex of the cat. Journal of Physiology (London) 500: 751–74.CrossRefGoogle ScholarPubMed
Brinley-Reed, M., Mascagni, F. and McDonald, A. J. (1995) Synaptology of prefrontal projections to the basolateral amygdala: an electron microscopic study in the rat. Neuroscience Letters 202: 45–8.CrossRefGoogle ScholarPubMed
Brodmann, K. (1905) Beitraege zur histologischen Lokalisation der Grosshirnrinde. III Mitteilung. Die Rindenfelder der niederen Affen. Journal für Psychologie und Neurologie 4: 177–226.Google Scholar
Brodmann, K. (1908) Beitraege zur histologischen Lokalisation der Grosshirnrinde. VI Mitteilung. Die Cortexgliederung des Menschen. Journal für Psychologie und Neurologie 10: 231–46.Google Scholar
Brown, M. W. and Aggleton, J. P. (2001) Recognition memory: what are the roles of the perirhinal cortex and hippocampus?Nature Reviews Neuroscience 2: 51–61.CrossRefGoogle ScholarPubMed
Brown, M. W. and Bashir, Z. I. (2002) Evidence concerning how neurons of the perirhinal cortex may affect familiarity discrimination. Philosophical Transactions of the Royal Society of LondonB357: 1083–95.CrossRefGoogle Scholar
Brown, M. W., Wilson, F. A. W. and Riches, I. P. (1987) Neuronal evidence that inferomedial temporal cortex is more important than hippocampus in certain processes underlying recognition memory. Brain Research 409: 158–67.CrossRefGoogle ScholarPubMed
Brown, R. E., Stevens, D. R. and Haas, H. L. (2001) The physiology of brain histamine. Progress in Neurobiology 63: 637–72.CrossRefGoogle ScholarPubMed
Buchanan, S. L., Thompson, R. H., Maxwell, B. L. and Powell, D. A. (1994) Efferent connections of the medial prefrontal cortex in the rabbit. Experimental Brain Research 100: 469–83.CrossRefGoogle ScholarPubMed
Budde, T., Munsch, T. and Pape, H. C. (1998) Distribution of L-type calcium channels in rat thalamic neurones. European Journal of Neuroscience 10: 586–97.CrossRefGoogle ScholarPubMed
Burette, F., Jay, T. M. and Laroche, S. (1997) Reversal of LTP in the hippocampal afferent fiber system to the prefrontal cortex in vivo with low-frequency patterns of stimulation that do not produce LTD. Journal of Neurophysiology 78: 1155–60.CrossRefGoogle Scholar
Burlhis, T.M and Aghajanian, G. K. (1987) Pacemaker potentials of serotonergic dorsal raphe neurons: contribution of a low-threshold C a2+ conductance. Synapse 1: 582–8.CrossRefGoogle Scholar
Burwell, R. D. (2001) Borders and cytoarchitecture of the perirhinal and postrhinal cortices in the rat. Journal of Comparative Neurology 437: 17–41.CrossRefGoogle ScholarPubMed
Burwell, R. D. and Witter, M. P. (2002) Basic anatomy of the parahippocampal region in monkeys and rats. In The Parahippocampal Region, ed. Witter, M. P. and Wouterlood, F., pp. 35–59. New York: Oxford University Press.Google Scholar
Buzsáki, G. (1986) Hippocampal sharp waves: their origin and significance. Brain Research 398: 242–52.CrossRefGoogle ScholarPubMed
Buzsáki, G. (1989) Two-stage model of memory trace formation: a role for ‘noisy’ brain states. Neuroscience 31: 551–70.CrossRefGoogle ScholarPubMed
Buzsáki, G. (1996) The hippocampo-neocortical dialogue. Cerebral Cortex 6: 81–92.CrossRefGoogle ScholarPubMed
Buzsáki, G. (2002) Theta oscillations in the hippocampus. Neuron 33: 325–40.CrossRefGoogle ScholarPubMed
Buzsáki, G. and Draguhn, A. (2004) Neuronal oscillations in cortical networks. Science 304: 1926–9.CrossRefGoogle ScholarPubMed
Buzsáki, G. and Eidelberg, E. (1982) Direct afferent excitation and long-term potentiation of hippocampal interneurons. Journal of Neurophysiology 48: 597–607.CrossRefGoogle ScholarPubMed
Buzsáki, G., Leung, L. and Vanderwolf, C. H. (1983) Cellular bases of hippocampal EEG in the behaving rat. Brain Research 6: 139–71.CrossRefGoogle Scholar
Buzsáki, G., Czopf, J., Kondákor, I. and Kellényi, L. (1986) Laminar distribution of hippocampal rhythmic slow activity (RSA) in the behaving rat: current-source density analysis, effects of urethane and atropine. Brain Research 365: 125–37.CrossRefGoogle ScholarPubMed
Buzsáki, G., Bickford, R. G., Ponomareff, G.et al. (1988) Nucleus basalis and thalamic control of neocortical activity in the freely moving rat. Journal of Neuroscience 8: 4007–26.CrossRefGoogle ScholarPubMed
Buzsáki, G., Kennedy, B., Solt, V. B. and Ziegler, M. (1991) Noradrenergic control of thalamic oscillations: the role of α-2 receptors. European Journal of Neuroscience 3: 222–9.CrossRefGoogle Scholar
Buzsáki, G., Horváth, Z., Urioste, R., Hetke, J. and Wise, K. (1992) High-frequency network oscillation in the hippocampus. Science 256: 1025–7.CrossRefGoogle ScholarPubMed
Cahill, L. and McGaugh, J. L. (1998) Mechanisms of emotional arousal and lasting declarative memory. Trends in Neurosciences 21: 294–9.CrossRefGoogle ScholarPubMed
Cahill, L., Babinsky, R., Markowitsch, H. J. and McGaugh, J. L. (1995) The amygdala and emotional memory. Nature 377: 295–6.CrossRefGoogle ScholarPubMed
Cahill, L., Haier, R. J., Fallon, J.et al. (1996) Amygdala activity at encoding correlated with long-term, free recall of emotional information. Proceedings of the National Academy of Sciences USA 93: 8016–21.CrossRefGoogle ScholarPubMed
Cahill, L., Weinberger, N. M., Roozendaal, B. and McGaugh, J. L. (1999) Is the amygdala a locus of ‘conditioned fear’? Some questions and caveats. Neuron 23: 227–8.CrossRefGoogle Scholar
Cantero, J. L., Atienza, M., Salas, R. M. and Dominguez-Marin, E. (2002) Effects of prolonged waking-auditory stimulation on electroencephalogram synchronization and cortical coherence during subsequent slow-wave sleep. Journal of Neuroscience 22: 4702–8.CrossRefGoogle ScholarPubMed
Cape, E. G. and Jones, B. E. (2000) Effects of glutamate agonist versus procaine microinjections into the basal forebrain cholinergic cell area upon gamma and theta EEG activity and sleep-wake state. European Journal of Neuroscience 12: 2166–84.CrossRefGoogle ScholarPubMed
Cardin, J. A., Palmer, L. A. and Contreras, D. (2005) Stimulus-dependent gamma (30–50 Hz) oscillations in simple and complex fast rhythmic bursting cells in primary visual cortex. Journal of Neuroscience 25: 5339–50.CrossRefGoogle ScholarPubMed
Carlsen, J. (1988) Immunocytochemical localization of glutamate decarboxylase in the rat basolateral amygdaloid nucleus, with special reference to GABAergic innervation of amygdalostriatal projection neurons. Journal of Comparative Neurology 273: 513–26.CrossRefGoogle ScholarPubMed
Carlsen, J. and Heimer, L. (1986) A correlated light and electron microscopic immunocytochemical study of cholinergic terminals and neurons in the rat amygdaloid body with special emphasis on the basolateral amygdaloid nucleus. Journal of Comparative Neurology 244: 121–36.CrossRefGoogle ScholarPubMed
Carlsen, J., Zaborszky, L. and Heimer, L. (1985) Cholinergic projections from the basal forebrain to the basolateral amygdaloid complex: a combined retrograde fluorescent and immunohistochemical study. Journal of Comparative Neurology 234: 155–67.CrossRefGoogle ScholarPubMed
Carmichael, S. T. and Price, J. L. (1995a) Limbic connections of the orbital and medial prefrontal cortex in macaque monkeys. Journal of Comparative Neurology 363: 615–41.CrossRefGoogle Scholar
Carmichael, S. T. and Price, J. L. (1995b) Sensory and premotor connections of the orbital and medial prefrontal cortex of macaque monkeys. Journal of Comparative Neurology 363: 642–64.CrossRefGoogle Scholar
Carr, C. E. and Konishi, M. (1990) A circuit for detection of interaural time differences in the brainstem of the barn owl. Journal of Neuroscience 10: 3227–46.CrossRefGoogle ScholarPubMed
Carr, D. B. and Sesack, S. R. (2000a) Dopamine terminals synapse on callosal projection neurons in the rat prefrontal cortex. Journal of Comparative Neurology 425: 275–83.3.0.CO;2-Z>CrossRefGoogle Scholar
Carr, D. B. and Sesack, S. R. (2000b) GABA-containing neurons in the rat ventral tegmental area project to the prefrontal cortex. Synapse 38: 114–23.3.0.CO;2-R>CrossRefGoogle Scholar
Carr, D. B. and Sesack, S. R. (2000c) Projections from the rat prefrontal cortex to the ventral tegmental area: target specificity in the synaptic associations with mesoaccumbens and mesocortical neurons. Journal of Neuroscience 20: 3864–73.CrossRefGoogle Scholar
Carskadon, M. A. and Dement, W. C. (2000) Normal human sleep: an overview. In Principles and Practice of Sleep Medicine, ed. Kryger, M. H., Roth, T. and Dement, W. C., pp. 15–25. Philadelphia, PA: W. B. Saunders.Google Scholar
Cassell, M. D. and Gray, T. S. (1989a) Morphology of peptide-immunoreactive neurons in the rat central nucleus of the amygdala. Journal of Comparative Neurology 281: 320–33.CrossRefGoogle Scholar
Cassell, M. D. and Gray, T. S. (1989b) The amygdala directly innervates adrenergic (C1) neurons in the ventrolateral medulla in the rat. Neuroscience Letters 97: 163–8.CrossRefGoogle Scholar
Cassell, M. D., Gray, T. S. and Kiss, J. Z. (1986) Neuronal architecture in the rat central nucleus of the amygdala: a cytological, hodological, and immunocytochemical study. Journal of Comparative Neurology 246: 478–99.CrossRefGoogle ScholarPubMed
Castaigne, P., Buge, A., Escourolle, R. and Mason, M. (1962) Ramollissement pédonculaire médian, tegmentothalamique avec ophthalmoplégie et hypersomnie. Revue Neurologique (Paris) 106: 357–67.Google Scholar
Castellano, C., Brioni, J. D., Nagahara, A. H. and McGaugh, J. L. (1989) Post-training systemic and intra-amygdala administration of the GABA-B agonist baclofen impairs retention. Behavioral and Neural Biology 52: 170–9.CrossRefGoogle ScholarPubMed
Castelo-Branco, M., Neuenschwander, S. and Singer, W. (1998) Synchronization of visual responses between the cortex, lateral geniculate nucleus, and retina in the anesthetized cat. Journal of Neuroscience 18: 6395–410.CrossRefGoogle ScholarPubMed
Castro-Alamancos, M. A. (2002a) Properties of primary sensory (lemniscal) synapses in the ventrobasal thalamus and the relay of high-frequency sensory inputs. Journal of Neurophysiology 87: 946–53.CrossRefGoogle Scholar
Castro-Alamancos, M. A. (2002b) Different temporal processing of sensory inputs in the rat thalamus during quiescent and information processing states in vivo. Journal of Physiology (London) 539: 567–78.CrossRefGoogle Scholar
Castro-Alamancos, M. A. and Calcagnotto, M. E. (2001) High-pass filtering of corticothalamic activity by neuromodulators released in the thalamus during arousal: in vitro and in vivo. Journal of Neurophysiology 85: 1489–97.CrossRefGoogle ScholarPubMed
Castro-Alamancos, M. A. and Connors, B. W. (1996) Cellular mechanisms of the augmenting response: short-term plasticity in a thalamocortical pathway. Journal of Neuroscience 16: 7742–56.CrossRefGoogle Scholar
Cauli, B., Audinat, E., Lambolez, B.et al. (1997) Molecular and physiological diversity of cortical nonpyramidal cells. Journal of Neuroscience 17: 3894–906.CrossRefGoogle ScholarPubMed
Cauller, L. J. and Kulics, A. T. (1988) A comparison of awake and sleeping cortical states by analysis of the somatosensory-evoked response of postcentral area 1 in rhesus monkey. Experimental Brain Research 72: 584–92.CrossRefGoogle ScholarPubMed
Cavada, C. and Reinoso-Suarez, F. (1985) Topographical organization of the cortical afferent connections of the prefrontal cortex in the cat. Journal of Comparative Neurology 242: 293–324.CrossRefGoogle ScholarPubMed
Cederbaum, J. M. and Aghajanian, G. K. (1978) Afferent projections to the rat locus coeruleus as determined by a retrograde tracing technique. Journal of Comparative Neurology 178: 1–16.CrossRefGoogle Scholar
Chagnac-Amitai, Y. and Connors, B. W. (1989) Synchronized excitation and inhibition driven by bursting neurons in neocortex. Journal of Neurophysiology 62: 1149–62.CrossRefGoogle ScholarPubMed
Chapman, P. F. and Bellavance, L. L. (1992) Induction of long-term potentiation in the basolateral amygdala does not depend on NMDA receptor activation. Synapse 11: 310–18.CrossRefGoogle Scholar
Chapman, P. F., Kairiss, E. W., Keenan, C. L. and Brown, T. H. (1990) Long-term synaptic potentiation in the amygdala. Synapse 6: 271–8.CrossRefGoogle ScholarPubMed
Chen, J. C. and Lang, E. J. (2003) Inhibitory control of rat lateral amygdaloid projection cells. Neuroscience 121: 155–66.CrossRefGoogle ScholarPubMed
Chen, W., Zhang, J. J., Hu, G. Y. and Wu, C. P. (1996) Electrophysiological and morphological properties of pyramidal and non-pyramidal neurons in the cat motor cortex in vitro. Neuroscience 73: 39–55.CrossRefGoogle Scholar
Chi, C. C. and Flynn, J. P. (1971) Neuroanatomic projections related to biting attack elicited from hypothalamus in cats. Brain Research 35: 49–66.CrossRefGoogle ScholarPubMed
Cho, K., Kemp, N., Noel, J., Aggleton, J. P., Brown, M. W. and Bashir, Z. I. (2000) A new form of long-term depression in the perirhinal cortex. Nature Neuroscience 3: 150–6.CrossRefGoogle ScholarPubMed
Cho, K., Aggleton, J. P., Brown, M. W. and Bashir, Z. I. (2001) An experimental test of the role of postsynaptic calcium levels in determining synaptic strength using perirhinal cortex of rat. Journal of Physiology (London) 532: 459–66.CrossRefGoogle ScholarPubMed
Chow, A., Farb, C., Nadal, M. S.et al. (1999) K+ channel expression distinguishes subpopulations of parvalbumin- and somatostatin-containing neocortical interneurons. Journal of Neuroscience 19: 9332–45.CrossRefGoogle ScholarPubMed
Chrobak, J. J. and Buzsáki, G. (1994) Selective activation of deep layers (V–VI) retrohippocampal cortical neurons during hippocampal sharp waves in the behaving rat. Journal of Neuroscience 14: 6160–70.CrossRefGoogle ScholarPubMed
Chrobak, J. J. and Buzsáki, G. (1996) High-frequency oscillations in the output networks of the hippocampal-entorhinal axis of the freely behaving rat. Journal of Neuroscience 16: 3056–66.CrossRefGoogle ScholarPubMed
Chrobak, J. J. and Buzsáki, G. (1998) Gamma oscillations in the entorhinal cortex of the freely behaving rat. Journal of Neuroscience 18: 388–98.CrossRefGoogle ScholarPubMed
Cicogna, P, Natale, V., Occhionero, M and Bosinelli, M. (2000) Slow wave sleep and REM sleep mentation. Sleep Research Online 3: 67–72.Google ScholarPubMed
Cirelli, C. and Tononi, G. (2000) Differential expression of plasticity-related genes in waking and sleep and their regulation by the noradrenergic system. Journal of Neuroscience 20: 9184–7.CrossRefGoogle ScholarPubMed
Cirelli, C. and Tononi, G. (2004) Locus coeruleus control of state-dependent gene expression. Journal of Neuroscience 24: 5410–19.CrossRefGoogle ScholarPubMed
Cirelli, C., Pompeiano, M. and Tononi, G. (1996) Neuronal gene expression in the waking state: a role for the locus coeruleus. Science 274: 1211–15.CrossRefGoogle ScholarPubMed
Cirelli, C., Gutierrez, C. M. and Tononi, G. (2004) Extensive and divergent effects of sleep and wakefulness on brain gene expression. Neuron 41: 35–43.CrossRefGoogle ScholarPubMed
Cirelli, C., Huber, R., Gopalakrishnan, A., Southard, T. L. and Tononi, G. (2005) Locus coeruleus control of slow-wave homeostasis. Journal of Neuroscience 25: 4503–11.CrossRefGoogle ScholarPubMed
Cissé, Y., Grenier, F., Timofeev, I. and Steriade, M. (2003) Electrophysiological properties and input-output organization of callosal neurons in cat association cortex. Journal of Neurophysiology 89: 1402–13.CrossRefGoogle ScholarPubMed
Cissé, Y., Crochet, S., Timofeev, I. and Steriade, M. (2004) Synaptic enhancement induced through callosal pathways in cat association cortex. Journal of Neurophysiology 92: 3221–32.CrossRefGoogle ScholarPubMed
Clugnet, M. C. and LeDoux, J. E. (1990) Synaptic plasticity in fear conditioning circuits: induction of LTP in the lateral nucleus of the amygdala by stimulation of the medial geniculate body. Journal of Neuroscience 10: 2818–24.CrossRefGoogle ScholarPubMed
Coenen, A. M. L. and Vendrik, A. J. H. (1972) Determination of the transfer ratio of cat's geniculate neurons through quasi-intracellular recordings and the relation with the level of alertness. Experimental Brain Research 14: 227–42.CrossRefGoogle ScholarPubMed
Colino, A. and Fernánde, Molina, A. (1986) Electrical activity generated in the subicular complex and basolateral amygdala of the rat. Neuroscience 19: 573–80.CrossRefGoogle ScholarPubMed
Collins, D. R. and Paré, D. (2000) Differential fear conditioning induces reciprocal changes in the sensory responses of lateral amygdala neurons to the CS(+) and CS(−). Learning and Memory 7: 97–103.CrossRefGoogle Scholar
Collins, D. R., Lang, E. J. and Paré, D. (1999) Spontaneous activity of the perirhinal cortex in behaving cats. Neuroscience 89: 1025–39.CrossRefGoogle ScholarPubMed
Collins, D. R., Pelletier, J. G. and Paré, D. (2001) Slow and fast (gamma) neuronal oscillations in the perirhinal cortex and lateral amygdala. Journal of Neurophysiology 85: 1661–72.CrossRefGoogle ScholarPubMed
Colombo, M. and Gross, C. G. (1994) Responses of inferior temporal cortex and hippocampal neurons during delayed matching to sample in monkeys (Macaca fascicularis). Behavioral Neuroscience 108: 443–55.CrossRefGoogle Scholar
Colonnier, M. (1966) The structural design of the neocortex. In Brain and Conscious Experience, ed. Eccles, J. C., pp. 1–23. New York: Springer.Google Scholar
Compte, A., Sanchez-Vives, M. V., McCormick, D. A. and Wang, X. J. (2003) Cellular and network mechanisms of slow oscillatory activity (<1 Hz) and wave propagations in a cortical network model. Journal of Neurophysiology 89: 2707–25.CrossRefGoogle Scholar
Condé, F., Audinat, E., Maire-Lepoivre, E. and Crépel, F. (1990) Afferent connections of the medial frontal cortex of the rat. A study using retrograde transport of fluorescent dyes. I. Thalamic afferents. Brain Research Bulletin 24: 341–54.CrossRefGoogle Scholar
Condé, F., E., M.-L., Audinat, E. and Crépel, F. (1995) Afferent connections of the medial frontal cortex of the rat. II. Cortical and subcortical afferents. Journal of Comparative Neurology 352: 567–93.CrossRefGoogle ScholarPubMed
Condorelli, D. F., Belluard, N., Trovato-Salinaro, A. and Mudo, I. (2000) Expression of Cx36 in mammalian neurons. Brain Research Reviews 32: 72–85.CrossRefGoogle ScholarPubMed
Connors, B. W. and Amitai, Y. (1995) Functions of local circuits in neocortex: synchrony and laminae. In The Cortical Neuron, ed. Gutnick, M. J. and Mody, I., pp. 123–40. New York: Oxford University Press.Google Scholar
Connors, B. W. and Gutnick, M. J. (1990) Intrinsic firing patterns of diverse neocortical neurons. Trends in Neurosciences 13: 99–104.CrossRefGoogle ScholarPubMed
Connors, B. W., Gutnick, M. J. and Prince, D. A. (1982) Electrophysiological properties of neocortical neurons in vitro. Journal of Neurophysiology 48: 1302–20.CrossRefGoogle ScholarPubMed
Connors, B. W., Malenka, R. and Silva, L. R. (1988) Two inhibitory postsynaptic potentials, and GAB AA and GABAB receptor-mediated responses in neocortex of rat and cat. Journal of Physiology (London) 406: 443–68.CrossRefGoogle Scholar
Contreras, D. and Llinás, R. (2001) Voltage-sensitive dye imaging of neocortical spatio-temporal dynamics to afferent activation frequency. Journal of Neuroscience 21: 9403–13.CrossRefGoogle Scholar
Contreras, D. and Palmer, L. (2003) Response to contrast of electrophysiologically defined cell classes in primary visual cortex. Journal of Neuroscience 23: 6936–45.CrossRefGoogle ScholarPubMed
Contreras, D. and Steriade, M. (1995) Cellular basis of EEG slow rhythms: a study of dynamic corticothalamic relationships. Journal of Neuroscience 15: 604–22.CrossRefGoogle ScholarPubMed
Contreras, D. and Steriade, M. (1996) Spindle oscillation: the role of corticothalamic feedback in a thalamically generated rhythm. Journal of Physiology (London) 490: 159–79.CrossRefGoogle Scholar
Contreras, D. and Steriade, M. (1997a) Synchronization of low-frequency rhythms in corticothalamic networks. Neuroscience 76: 11–24.CrossRefGoogle Scholar
Contreras, D. and Steriade, M. (1997b) State-dependent fluctuations of low-frequency rhythms in corticothalamic networks. Neuroscience 76: 25–38.CrossRefGoogle Scholar
Contreras, D., Curró Dossi, R. and Steriade, M. (1993) Electrophysiological properties of cat reticular neurones in vivo. Journal of Physiology (London) 470: 273–94.CrossRefGoogle ScholarPubMed
Contreras, D., Destexhe, A., Sejnowski, T. J. and Steriade, M. (1996a) Control of spatiotemporal coherence of a thalamic oscillation by corticothalamic feedback. Science 274: 771–4.CrossRefGoogle Scholar
Contreras, D., Timofeev, I. and Steriade, M. (1996b) Mechanisms of long-lasting hyperpolarizations underlying slow sleep oscillations in cat corticothalamic networks. Journal of Physiology (London) 494: 251–64.CrossRefGoogle Scholar
Contreras, D., Destexhe, A. and Steriade, M. (1997a) Intracellular and computational characterization of the intracortical inhibitory control of synchronized thalamic inputs in vivo. Journal of Neurophysiology 78: 335–50.CrossRefGoogle Scholar
Contreras, D., Destexhe, A., Sejnowski, T. J. and Steriade, M. (1997b) Spatiotemporal patterns of spindle oscillations in cortex and thalamus. Journal of Neuroscience 17: 1179–96.CrossRefGoogle Scholar
Contreras, D., Dürmüller, N. and Steriade, M. (1997c) Absence of a prevalent laminar distribution of IPSPs in association cortical neurons of cat. Journal of Neurophysiology 78: 2742–53.CrossRefGoogle Scholar
Cossart, R., Aronov, D. and Yuste, R. (2003) Attractor dynamics of network UP states in the neocortex. Nature 423: 283–8.CrossRefGoogle ScholarPubMed
Cox, C. L. and Sherman, S. M. (2000) Control of dendritic outputs of inhibitory interneurons in the lateral geniculate nucleus. Neuron 27: 597–610.CrossRefGoogle ScholarPubMed
Cox, C. L., Huguenard, J. R. and Prince, D. A. (1996) Heterogenous axonal arborizations of rat thalamic reticular neurons in the ventrobasal nucleus. Journal of Comparative Neurology 366: 416–30.3.0.CO;2-7>CrossRefGoogle Scholar
Cox, C. L., Huguenard, J. R. and Prince, D. A. (1997) Nucleus reticularis neurons mediate diverse inhibitory effects in the thalamus. Proceedings of the National Academy of Sciences USA 94: 8854–9.CrossRefGoogle ScholarPubMed
Cox, C. L., Zhou, Q. and Sherman, S. M. (1998) Glutamate locally activates dendritic outputs of thalamic interneurons. Nature 394: 478–82.CrossRefGoogle ScholarPubMed
Cox, C. L., Reichova, I. and Sherman, S. M. (2003) Functional synaptic contacts by intranuclear axon collaterals of thalamic relay neurons. Journal of Neuroscience 23: 7642–6.CrossRefGoogle ScholarPubMed
Cox, G. E., Jordan, D., Moruzzi, P.et al. (1986) Amygdaloid influences on brain-stem neurones in the rabbit. Journal of Physiology (London) 381: 135–48.CrossRefGoogle ScholarPubMed
Crabtree, J. W., Collingridge, G. L. and Isaac, J. T. R. (1998) A new intrathalamic pathway linking modality-related nuclei in the dorsal thalamus. Nature Neuroscience 1: 389–94.CrossRefGoogle ScholarPubMed
Crick, F. (1984) The function of the thalamic reticular complex: the searchlight hypothesis. Proceedings of the National Academy of Sciences USA 81: 4586–90.CrossRefGoogle ScholarPubMed
Crick, F. (1994) The Astonishing Hypothesis. New York: Charles Scribner's Sons.Google Scholar
Crill, W. E. (1996) Persistent sodium current in mammalian central neurones. Annual Review of Physiology 58: 349–62.CrossRefGoogle Scholar
Crochet, S., Fuentealba, P., Timofeev, I. and Steriade, M. (2004) Selective amplification of neocortical neuronal output by fast prepotentialsin vivo. Cerebral Cortex 14: 1110–21.CrossRefGoogle ScholarPubMed
Crosby, E. C. and Humphrey, T. (1941) Studies of the vertebrate telencephalon. II. The nuclear pattern of the anterior olfactory nucleus, tuberculum olfactorum and the amygdaloid complex in adult man. Journal of Comparative Neurology 74: 309–52.CrossRefGoogle Scholar
Crow, T. J. (1997) Schizophrenia as failure of hemispheric dominance for language. Trends in Neuroscience 20: 339–43.Google ScholarPubMed
Crunelli, V. and Leresche, N. (2002) Childhood absence epilepsy: genes, channels, neurons and networks. Nature Reviews Neuroscience 3: 371–82.CrossRefGoogle ScholarPubMed
Crunelli, V., Haby, M., Jassik-Gerschenfeld, D., Leresche, N. and Pirchio, M. (1988) C l− and K+-dependent inhibitory postsynaptic potentials evoked by interneurons of the rat lateral geniculate nucleus. Journal of Physiology (London) 399: 153–76.CrossRefGoogle Scholar
Crutcher, M. D., Branch, M. H., DeLong, M. R. and Georgopoulos, A. P. (1980) Activity of zona incerta neurons in the behaving primate. Society for Neuroscience Abstracts 6: 676.Google Scholar
Csicsvari, J., Hirase, H., Czurkó, A., Mamiya, A. and Buzsáki, G. (1999) Oscillatory coupling of hippocampal pyramidal cells and interneurons in the behaving cat. Journal of Neuroscience 19: 274–87.CrossRefGoogle Scholar
Cunningham, E. T. and LeVay, S. (1986) Laminar and synaptic organization of the projection from the thalamic nucleus centralis to primary visual cortex in the cat. Journal of Comparative Neurology 254: 65–77.CrossRefGoogle ScholarPubMed
Cunningham, M. O., Whittington, M. A., Bibbig, A.et al. (2004) A role for fast rhythmic bursting neurons in cortical gamma oscillations in vitro. Proceedings of the National Academy of Sciences USA 101: 7152–7.CrossRefGoogle ScholarPubMed
Curró Dossi, R., Paré, D. and Steriade, M. (1991) Short-lasting nicotinic and long-lasting muscarinic depolarizing responses of thalamocortical neurons to stimulation of mesopontine cholinergic nuclei. Journal of Neurophysiology 65: 393–406.CrossRefGoogle ScholarPubMed
Curró Dossi, R., Nuñez, A. and Steriade, M. (1992a) Electrophysiology of a slow (0.5–4 Hz) intrinsic oscillation of cat thalamocortical neurones in vivo. Journal of Physiology (London) 447: 215–34.CrossRefGoogle Scholar
Curró Dossi, R., Paré, D. and Steriade, M. (1992b) Various types of inhibitory postsynaptic potentials in anterior thalamic cells are differentially altered by stimulation of laterodorsal tegmental cholinergic nucleus. Neuroscience 47: 279–89.CrossRefGoogle Scholar
Damasio, A. R. (1996) The somatic marker hypothesis and the possible functions of the prefrontal cortex. Philosophical Transactions of the Royal Society of LondonB351: 1413–20.CrossRefGoogle ScholarPubMed
Damasio, A. R. and Maurer, R. G. (1978) A neurological model for childhood autism. Archives of Neurology 35: 777–86.CrossRefGoogle ScholarPubMed
Danober, L. and Pape, H. C. (1998) Mechanisms and functional significance of a slow inhibitory potential in neurons of the lateral amygdala. European Journal of Neuroscience 10: 853–67.CrossRefGoogle ScholarPubMed
Datta, S., Curró Dossi, R., Paré, D., Oakson, G. and Steriade, M. (1991) Substantia nigra reticulata neurons during sleep-waking states: relation with ponto-geniculo-occipital waves. Brain Research 566: 344–7.CrossRefGoogle ScholarPubMed
Davenne, D. and Adrien, J. (1984) Suppression of PGO waves in the kitten: anatomical effects on the lateral geniculate nucleus. Neuroscience Letters 45: 33–8.CrossRefGoogle ScholarPubMed
Davis, M. (1992) The role of the amygdala in fear and anxiety. Annual Review of Neuroscience 15: 353–75.CrossRefGoogle ScholarPubMed
Davis, M. (2000) The role of the amygdala in conditioned and unconditioned fear and anxiety. In The Amygdala: A Functional Analysis, ed. Aggleton, J. P., pp. 213–87. Oxford: Oxford University Press.Google Scholar
Dawson, T., Bredt, D., Fotuhi, M., Hwang, P. and Snyder, S. (1991) Nitric oxide synthase and neuronal NADPH diaphorase are identical in brain and peripheral tissues. Proceedings of the National Academy of Sciences USA 88: 7797–801.CrossRefGoogle ScholarPubMed
Deacon, T. W., Eichenbaum, H., Rosenberg, P. and Eckman, K. W. (1983) Afferent connections of the perirhinal cortex in the rat. Journal of Comparative Neurology 220: 168–90.CrossRefGoogle ScholarPubMed
DeFelipe, J. and Fariñas, I. (1992) The pyramidal neuron of the cerebral cortex: morphological and chemical characteristics of the synaptic inputs. Progress in Neurobiology 39: 563–607.CrossRefGoogle ScholarPubMed
DeFelipe, J. and Jones, E. G. (1992) High-resolution light and electron microscopy immunocytochemistry of colocalized GABA and calbindin D-28k in somata and double bouquet cell axons in the monkey sensory-motor cortex. European Journal of Neuroscience 4: 46–60.CrossRefGoogle Scholar
DeFelipe, J., Elston, G. N., Fujita, I.et al. (2002) Neocortical circuits: evolutionary aspects and specificity versus non-specificity of synaptic connections. Remarks, main conclusions and general comments and discussion. Journal of Neurocytology 31: 387–416.CrossRefGoogle ScholarPubMed
Gennaro, L., Ferrara, M. and Bertini, M. (2001) The boundary between wakefulness and sleep: quantitative electroencephalographic changes during the sleep onset period. Neuroscience 107: 1–11.CrossRefGoogle ScholarPubMed
Demaurex, N., Lew, D. P. and Krause, K.-H. (1992) Cyclopiazonic acid depletes intracellular Ca2+ stores and activates an influx pathway for divalent cations in HL-60 cells. Journal of Biological Chemistry 267: 2318–24.Google ScholarPubMed
Dement, W. C., Ferguson, J., Cohen, H. and Barchas, J. (1969) Non-chemical methods and data using a biochemical model: the REM quanta. In Psychochemical Research in Man: Methods, Strategy and Theory, ed. Mandell, A. and Mandell, M. P., pp. 275–325. New York: Academic Press.Google Scholar
Denning, K. S. and Reinagel, P. (2005) Visual control of burst priming in the anesthetized lateral geniculate nucleus. Journal of Neuroscience 25: 3531–8.CrossRefGoogle ScholarPubMed
Oca, B. M., Cola, J. P., Maren, S. and Fanselow, M. S. (1998) Distinct regions of the periaqueductal gray are involved in the acquisition and expression if defensive responses. Journal of Neuroscience 18: 3426–32.CrossRefGoogle ScholarPubMed
Olmos, J. S. (1990) Amygdala. In The Human Nervous System, ed. Paxinos, G., pp. 583–708. New York: Academic Press.Google Scholar
Descarries, L., Watkins, K. C., and Lapierre, Y. (1977) Noradrenergic axon terminals in the cerebral cortex of rat. III. Topometric ultrastructural analysis. Brain Research 133: 197–222.CrossRefGoogle ScholarPubMed
Descarries, L., Gisiger, V. and Steriade, M. (1997) Diffuse transmission by acetylcholine in the CNS. Progress in Neurobiology 53: 603–25.CrossRefGoogle ScholarPubMed
Descarries, L., Mechawar, N., Aznavour, N. and Watkins, K. C. (2004) Structural determinants of the roles of acetylcholine in cerebral cortex. In Acetylcholine in the Cerebral Cortex (Progress in Brain Research, vol. 145), ed., Descarries, L., Krnjevic, K. and Steriade, M., pp. 45–58. Amsterdam: Elsevier.Google Scholar
Deschênes, M. (1981) Dendritic spikes induced in fast pyramidal tract neurons by thalamic stimulation. Experimental Brain Research 43: 304–8.Google ScholarPubMed
Deschênes, M. and Hu, B. (1990) Electrophysiology and pharmacology of the corticothalamic input to lateral thalamic nuclei: an intracellular study in the cat. European Journal of Neuroscience 2: 140–52.CrossRefGoogle Scholar
Deschênes, M. and Steriade, M. (1988) The neuronal mechanism of thalamic PGO waves. In Cellular Thalamic Mechanisms, ed. Bentivoglio, M., Macchi, G. and Spreafico, R., pp. 197–206. Amsterdam: Elsevier.Google Scholar
Deschênes, M., Roy, J. P. and Steriade, M. (1982) Thalamic bursting mechanism: an inward slow current revealed by membrane hyperpolarization. Brain Research 239: 289–93.CrossRefGoogle ScholarPubMed
Deschênes, M., Paradis, M., Roy, J. P. and Steriade, M. (1984) Electrophysiology of neurons of lateral thalamic nuclei in cat: resting properties and burst discharges. Journal of Neurophysiology 51: 1196–219.CrossRefGoogle ScholarPubMed
Deschênes, M., Madariaga-Domich, A. and Steriade, M. (1985) Dendrodendritic synapses in cat reticularis thalami nucleus, a structural basis for thalamic spindle synchronization. Brain Research 334: 169–71.Google ScholarPubMed
Deschênes, M., Veinante, P. and Zhang, Z. W. (1998).The organization of corticothalamic projections: reciprocity versus parity. Brain Research Reviews 28: 286–308.CrossRefGoogle ScholarPubMed
Desmedt, J. E. and Tomberg, C. (1994) Transient phase-locking of 40 Hz electrical oscillations in prefrontal and parietal human cortex reflects the process of conscious somatic perception. Neuroscience Letters 168: 126–9.CrossRefGoogle ScholarPubMed
Desmedt, J. E., Nguyen, T. H. and Bourget, M. (1983) The cognitive P40, N60 and P100 components of somatosensory evoked potentials and the earliest signs of sensory processing in man. Electroencephalography and Clinical Neurophysiology 56: 272–82.CrossRefGoogle ScholarPubMed
Destexhe, A. and Paré, D. (1999) Impact of network activity on the integrative properties of neocortical pyramidal neurons in vivo. Journal of Neurophysiology 81: 1531–47.CrossRefGoogle ScholarPubMed
Destexhe, A. and Paré, D. (2000) A combined computational and intracellular study of correlated synaptic bombardment in neocortical pyramidal neurons in vivo. Neurocomputing 32–3: 113–19.CrossRefGoogle Scholar
Destexhe, A., Contreras, D., Sejnowski, T. J. and Steriade, M. (1994a) A model of spindle rhythmicity in the isolated thalamic reticular nucleus. Journal of Neurophysiology 72: 803–18.CrossRefGoogle Scholar
Destexhe, A., Contreras, D., Sejnowski, T. J. and Steriade, M. (1994b) Modeling the control of reticular thalamic oscillations by neuromodulators. NeuroReport 5: 2217–20.CrossRefGoogle Scholar
Destexhe, A., Contreras, D., Steriade, M., Sejnowski, T. J. and Huguenard, J. R. (1996) In vivo, in vitro and computational analysis of dendritic calcium currents in thalamic reticular neurons. Journal of Neuroscience 16: 169–85.CrossRefGoogle ScholarPubMed
Destexhe, A., Contreras, D. and Steriade, M. (1999) Spatiotemporal analysis of local field potentials and unit discharges in cat cerebral cortex during natural wake and sleep states. Journal of Neuroscience 19: 4595–608.CrossRefGoogle ScholarPubMed
Destexhe, A., Contreras, D. and Steriade, M. (2001) LTS cells in cerebral cortex and their role in generating spike-and-wave oscillations. Neurocomputing 38–40: 555–63.CrossRefGoogle Scholar
Destexhe, A., Rudolph, M. and Paré, D. (2003) The high-conductance state of neocortical neurons in vivo. Nature Reviews Neuroscience 4: 739–51.CrossRefGoogle ScholarPubMed
Détári, L., Rasmusson, D. D. and Semba, K. (1997) Phasic relationship between the activity of basal forebrain neurons and cortical EEG in urethane-anesthetized rat. Brain Research 759: 112–21.CrossRefGoogle ScholarPubMed
Deuchars, J. and Thomson, A. M. (1995a) Single axon IPSPs elicited by a sparsely spiny interneuron in rat neocortex. Neuroscience 65: 935–42.CrossRefGoogle Scholar
Deuchars, J. and Thomson, A. M. (1995b) Innervation of burst firing spiny interneurons by pyramidal cells in deep layers of rat somatomotor cortex: paired intracellular recordings with biocytin filling. Neuroscience 69: 739–55.CrossRefGoogle Scholar
Dickinson, A. H., Mesches, M. H., Coleman, K. and McGaugh, J. L. (1993) Bicuculline administered into the amygdala blocks benzodiazepine-induced amnesia. Behavioral and Neural Biology 60: 1–4.CrossRefGoogle Scholar
Dickson, C. T., Biella, G. and Curtis, M. (2003) Slow periodic events and their transition to gamma oscillations in the entorhinal cortex of the isolated guinea pig brain. Journal of Neurophysiology 90: 39–46.CrossRefGoogle ScholarPubMed
Di Pasquale, E., Keegan, K. D. and Noebels, J. L. (1997) Increased excitability and inward rectification in layer V cortical pyramidal neurons in the epileptic mutant mouse Stargazer. Journal of Neurophysiology 77: 621–31.CrossRefGoogle Scholar
Dolan, R. J., Bench, C. J., Brown, R. G.et al. (1992) Regional cerebral blood flow abnormalities in depressed patients with cognitive impairment. Journal of Neurology, Neurosurgery and Psychiatry 55: 768–73.CrossRefGoogle ScholarPubMed
Dolmetsch, R. E., Pajvani, U., Fife, K., Spotts, J. M. and Greenberg, M. E. (2001) Signaling to the nucleus by an L-type calcium channel-calmodulin complex through the MAP kinase pathway. Science 294: 333–9.CrossRefGoogle ScholarPubMed
Domich, L., Oakson, G. and Steriade, M. (1986) Thalamic burst patterns in the naturally sleeping cat: a comparison between cortically projecting and reticularis neurones. Journal of Physiology (London) 379: 429–49.CrossRefGoogle ScholarPubMed
Douglas, R. J. and Martin, K. (1991) A functional microcircuit for cat visual cortex. Journal of Physiology (London) 440: 735–69.CrossRefGoogle ScholarPubMed
Douglas, R. J., Koch, C., Mahowald, M., Martin, K. A. C. and Suarez, H. H. (1995) Recurrent excitation in neocortical circuits. Science 269: 981–5.CrossRefGoogle ScholarPubMed
Doyère, V., Burette, F., Negro, C. R. and Laroche, S. (1993) Long-term potentiation of hippocampal afferents and efferents to prefrontal cortex: implications for associative learning. Neuropsychologia 31: 1031–53.CrossRefGoogle ScholarPubMed
Drevets, W. C. and Reichle, M. E. (1992) Neuroanatomical circuits in depression: implication for treatment mechanisms. Psychopharmacological Bulletin 28: 261–74.Google Scholar
Dringenberg, H. C. and Olmstead, M. C. (2003) Integrated contributions of basal forebrain and thalamus to neocortical activation elicited by pedunculopontine tegmental stimulation in urethane-anesthetized rats. Neuroscience 119: 839–53.CrossRefGoogle ScholarPubMed
Dringenberg, H. C. and Vanderwolf, C. H. (1996) Cholinergic activation of the electrocorticogram: an amygdaloid activating system. Experimental Brain Research 108: 285–96.CrossRefGoogle ScholarPubMed
Dumont, É.C., Martina, M., Samson, R. D., Drolet, G. and Paré, D. (2002) Physiological properties of central amygdala neurons: species differences. European Journal of Neuroscience 15: 545–52.CrossRefGoogle ScholarPubMed
Dumont, S. and Dell, P. (1960) Facilitation réticulaire des mécanismes visuels corticaux. Electroencephalography and Clinical Neurophysiology 12: 769–96.CrossRefGoogle Scholar
Duncan, C. P. (1949) The retroactive effect of electroshock on learning. Journal of Comparative Physiology and Psychology 42: 32–44.CrossRefGoogle ScholarPubMed
Duque, A., Balatoni, B., Détári, L. and Zaborsky, L. (2000) EEG correlation of the discharge properties of identified neurons in the basal forebrain. Journal of Neurophysiology 84: 1627–35.CrossRefGoogle ScholarPubMed
Durand, G. M., Kovalchuk, Y. and Konnerth, A. (1996) Long-term potentiation and functional synapse induction in developing hippocampus. Nature 381: 71–5.CrossRefGoogle ScholarPubMed
Eccles, J. C. (1961) Chairman's opening remarks. In The Nature of Sleep, ed. Wolstenholme, G. E. W. and O'Connor, M., pp. 1–3. London: Churchill.Google Scholar
Eckhorn, R., Bauer, R., Jordan, W.et al. (1988) Coherent oscillations: a mechanism of feature linking in the visual cortex?Biological Cybernetics 60: 121–30.CrossRefGoogle ScholarPubMed
Egan, T. M. and North, R. A. (1985) Acetylcholine acts on m2-muscarinic receptors to excite rat locus coeruleus neurones. British Journal of Pharmacology 85: 733–5.CrossRefGoogle ScholarPubMed
Ehlers, C., Hendricksen, S. J., Wang, M.et al. (1983) Corticotropin releasing factor produces increases in brain excitability and convulsive seizures in rats. Brain Research 278: 332–6.CrossRefGoogle ScholarPubMed
Eichenbaum, H. (2002) Memory representations in the parahippocampal region. In The Parahippocampal Region, ed. Witter, M. P. and Wouterlood, F., pp. 165–84. Oxford: Oxford University Press.Google Scholar
Eichenbaum, H., Schoenbaum, G., Young, B. and Bunsey, M. (1996) Functional organization of the hippocampal memory system. Proceedings of the National Academy of Sciences USA 93: 13500–7.CrossRefGoogle ScholarPubMed
El Mansari, M., Sakai, K. and Jouvet, M. (1989) Unitary characteristics of presumptive cholinergic tegmental neurons during the sleep-waking cycle in freely moving cats. Experimental Brain Research 76: 519–29.CrossRefGoogle ScholarPubMed
Emerson, R. G., Sgro, J. A., Pedley, T. A. and Hauser, A. (1988) State-dependent changes in the N20 component of the median nerve somatosensory evoked potential. Neurology 38: 64–7.CrossRefGoogle ScholarPubMed
Emptage, N., Bliss, T. and Fine, A. (1999) Single synaptic events evoke NMDA receptor-mediated release of calcium from internal stores in hippocampal dendritic spines. Neuron 22: 115–24.CrossRefGoogle ScholarPubMed
Endo, K., Araki, T. and Ito, K. (1977) Short latency EPSPs and incrementing PSPs of pyramidal tract cells evoked by stimulation of the nucleus centralis lateralis of the thalamus. Brain Research 132: 541–6.CrossRefGoogle ScholarPubMed
Engel, A. K., König, P., Kreiter, A. K. and Singer, W. (1991).Interhemispheric synchronization of oscillatory neuronal responses in cat visual cortex. Science 252: 1177–9.CrossRefGoogle ScholarPubMed
Engel, A. K., Fries, P and Singer, W. (2001) Dynamic predictions: oscillations and synchrony in top-down processing. Nature Reviews Neuroscience 2: 704–16.CrossRefGoogle ScholarPubMed
Esclapez, M., Tillakaratne, N., Tobin, A. J. and Houser, C. R. (1993) Comparative localization of two forms of glutamic acid decarboxylase with nonradioactive in situ hybridization methods. Journal of Comparative Neurology 331: 339–62.CrossRefGoogle ScholarPubMed
Evarts, E. V. (1964) Temporal patterns of discharge of pyramidal tract neurons during sleep and waking. Journal of Neurophysiology 27: 152–71.CrossRefGoogle ScholarPubMed
Evarts, E. V. (1965) Relation of discharge frequency to conduction velocity in pyramidal tract neurons. Journal of Neurophysiology 28: 216–28.CrossRefGoogle ScholarPubMed
Faber, E. and Sah, P. (2002) Physiological role of calcium-activated potassium currents in the rat lateral amygdala. Journal of Neuroscience 22: 1618–28.CrossRefGoogle ScholarPubMed
Faber, E. S. and Sah, P. (2003) Calcium-activated K+ (BK) channel inactivation contributes to spike broadening during repetitive firing in rat lateral amygdala neurons. Journal of Physiology (London) 552 (2): 482–97.CrossRefGoogle Scholar
Faber, E., Callister, R. J. and Sah, P. (2001) Morphological and electrophysiological properties of principal neurons in the rat lateral amygdala in vitro. Journal of Neurophysiology 85: 714–23.CrossRefGoogle ScholarPubMed
Façon, E., Steriade, M. and Wertheimer, N. (1958) Hypersomnie prolongée engendrée par des lésions bilatérales du système activateur médial: le syndrome thrombotique de la bifurcation du tronc basilaire. Revue Neurologique (Paris) 98: 117–33.Google Scholar
Fahy, F. L., Riches, I. P. and Brown, M. W. (1993) Neuronal activity related to visual recognition memory: long-term memory and the encoding of recency and familiarity information in the primate anterior and medial inferior temporal cortex and rhinal cortex. Experimental Brain Research 96: 457–72.CrossRefGoogle ScholarPubMed
Fallon, J. H. and Ciofi, P. (1992) Distribution of monoamines with the amygdala. In The Amygdala, ed. Aggleton, J. P., pp. 97–114. New York: Wiley-Liss.Google ScholarPubMed
Fallon, J. H., Koziell, D. A. and Moore, R. Y. (1978) Catecholamine innervation of the basal forebrain. II. Amygdala, suprarhinal cortex and entorhinal cortex. Journal of Comparative Neurology 180: 509–32.CrossRefGoogle ScholarPubMed
Falls, W. A., Miserendino, M. J. D. and Davis, M. (1992) Extinction of fear-potentiated startle: blockade by infusion of an NMDA antagonist into the amygdala. Journal of Neuroscience 12: 854–63.CrossRefGoogle ScholarPubMed
Fanselow, M. S. and Kim, J. J. (1994) Acquisition of contextual Pavlovian fear conditioning is blocked by application of an NMDA receptor antagonist D,L-2-amino-5-phosphonovaleric acid to the basolateral amygdala. Behavioral Neuroscience 108: 210–12.CrossRefGoogle ScholarPubMed
Fanselow, M. S. and LeDoux, J. E. (1999) Why we think plasticity underlying Pavlovian fear conditioning occurs in the basolateral amygdala. Neuron 23: 229–32.CrossRefGoogle ScholarPubMed
Fanselow, M. S., Kim, J. J., Yipp, J. and Oca, B. (1994) Differential effects of the N-methyl-D-aspartate antagonist DL-2-amino-5-phosphonovalerate on acquisition of fear of auditory and contextual cues. Behavioral Neuroscience 108: 235–40.CrossRefGoogle ScholarPubMed
Farb, C. R. and LeDoux, J. E. (1999) Afferents from rat temporal cortex synapse on lateral amygdala neurons that express NMDA and AMPA receptors. Synapse 33: 218–29.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Farmer, S. F. (1998) Rhythmicity, synchronization and binding in human primate cortex. Journal of Physiology (London) 509: 3–14.CrossRefGoogle Scholar
Faulkner, B. and Brown, T. H. (1999) Morphology and physiology of neurons in the rat perirhinal-lateral amygdala area. Journal of Comparative Neurology 411: 613–42.3.0.CO;2-U>CrossRefGoogle ScholarPubMed
Feinberg, I. and Campbell, I. G. (1993) Ketamine administration during waking increases delta EEG intensity in rat sleep. Neuropharmacology 9: 41–8.Google ScholarPubMed
Fellous, J. M., Rudolph, M., Destexhe, A. and Sejnowski, T. J. (2003) Synaptic background noise controls the input/output characteristics of single cells in an in vitro model of in vivo activity. Neuroscience 122: 811–29.CrossRefGoogle Scholar
Ferrara, M., Gennaro, L. and Bertini, M. (1999) Selective slow-wave sleep (SWS) deprivation and SWS rebound: do we need a fixed amount per night?Sleep Research Online 2: 15–19.Google Scholar
Ferrara, M., Gennaro, L., Curcio, G.et al. (2002) Regional differences of the human sleep electroencephalogram in response to selective slow-wave sleep deprivation. Cerebral Cortex 12: 737–48.CrossRefGoogle ScholarPubMed
Ferry, B. and McGaugh, J. L. (1999) Clenbuterol administration into the basolateral amygdala post-training enhances retention in an inhibitory avoidance task. Neurobiology of Learning and Memory 72: 8–12.CrossRefGoogle Scholar
Ferry, B., Magistretti, P. J. and Pralong, E. (1997) Noradrenaline modulates glutamate-mediated neurotransmission in the rat basolateral amygdala in vitro. European Journal of Neuroscience 9: 1356–64.CrossRefGoogle ScholarPubMed
Feshchenko, V. A. and Chilingaryan, L. I. (1990) Dependence of electrical activity of the amygdaloid complex on level of motivation and emotional state of the dog. Neuroscience Behavioral Physiology 20: 506–13.CrossRefGoogle ScholarPubMed
Finch, D. M., Wong, E. E., Derian, E. L., Chen, X. H., Nowlin-Finch, N. L. and Brothers, L. A. (1986) Neurophysiology of limbic system pathways in the rat: projections from the amygdala to the entorhinal cortex. Brain Research 370: 273–84.CrossRefGoogle ScholarPubMed
Finch, D. M., Tan, A. M. and Isokawa-Akesson, M. (1988) Feedforward inhibition of the rat entorhinal cortex and subicular complex. Journal of Neuroscience 8: 2213–26.CrossRefGoogle ScholarPubMed
Finelli, L. A., Baumann, H., Borbély, A. and Achermann, P. (2000) Dual electroencephalogram markers of human sleep homeostasis: correlation between theta activity in waking and slow-wave activity in sleep. Neuroscience 101: 523–9.CrossRefGoogle ScholarPubMed
Fiset, P., Paus, T., Daloze, T.et al. (1999) Brain mechanisms of propofol-induced loss of consciousness in humans: a positron emission tomography study. Journal of Neuroscience 19: 5506–13.CrossRefGoogle Scholar
Foehring, R. C. and Waters, R. S. (1991) Contributions of low-threshold calcium current and anomalous rectifier (Ih) to slow depolarizations underlying burst firing in human neocortical neurons in vitro. Neuroscience Letters 124: 17–21.CrossRefGoogle ScholarPubMed
Foehring, R. C., Schwindt, P. C. and Crill, W. E. (1989) Norepinephrine selectively reduces slow Ca2+- and Na+ -mediated K+ currents in cat neocortical neurons. Journal of Neurophysiology 61: 245–256.CrossRefGoogle ScholarPubMed
Fonnum, S., Strom-Mathisen, J. and Divac, I. (1981) Biochemical evidence for glutamate as neurotransmitter in corticostriatal and corticothalamic fibres in rat brain. Neuroscience 6: 863–73.CrossRefGoogle ScholarPubMed
Fontanini, A., Spano, P and Bower, J. M. (2003) Ketamine-xylazine-induced slow (<1.5 Hz) oscillations in the rat piriform (olfactory) cortex are functionally correlated with respiration. Journal of Neuroscience 23: 7993–8001.CrossRefGoogle ScholarPubMed
Foote, S. L., Bloom, F. E., and Aston-Jones, G. (1983) Nucleus locus ceruleus: new evidence of anatomical and physiological specificity. Physiological Reviews 63: 844–914.CrossRefGoogle ScholarPubMed
Ford, B., Holmes, C. J., Mainville, L. and Jones, B. E. (1995) GABAergic neurons in the rat pontomesencephalic tegmentum: codistribution with cholinergic and other tegmental neurons projecting to the posterior lateral hypothalamus. Journal of Comparative Neurology 363: 177–96.CrossRefGoogle ScholarPubMed
Forslid, A., Andersson, B. and Johansson, S. (1986) Observations on normal EEG activity in different brain regions of the unrestrained swine. Acta Physiologica Scandinavica 128: 389–96.CrossRefGoogle ScholarPubMed
Fort, P., Khateb, A., Pegna, A., Jones, B. E. and Mühlethaler, M. (1995) Noradrenergic modulation of cholinergic nucleus basalis neurons demonstrated by in vitro pharmacological and immunohistochemical evidence in the guinea pig brain. European Journal of Neuroscience 7: 1502–11.CrossRefGoogle ScholarPubMed
Fortin, M., Asselin, M. C., Gould, P. V. and Parent, A. (1998) Calretinin-immunoreactive neurons in the human thalamus. Neuroscience 84: 537–48.CrossRefGoogle ScholarPubMed
Fosse, R., Stickgold, R. and Hobson, J. A. (2004) Thinking and hallucinating: reciprocal changes in sleep. Psychophysiology 41: 298–305.CrossRefGoogle ScholarPubMed
Foster, J. A. (1980) Intracortical origin of recruiting responses in the cat neocortex. Electroencephalography and Clinical Neurophysiology 48: 639–53.CrossRefGoogle Scholar
Foulkes, D. (1962) Dream reports from different stages of sleep. Journal of Abnormal and Social Psychology 65: 14–25.CrossRefGoogle Scholar
Foulkes, D. and Schmidt, M. (1983) Temporal sequence and unit composition in dream reports from different stages of sleep. Sleep 6: 265–80.CrossRefGoogle Scholar
Fournier, G. N., Materi, L. M., Semba, K. and Rasmusson, D. D. (2004) Cortical acetylcholine release and electroencephalogram activation evoked by ionotropic glutamate receptor agonists in the rat basal forebrain. Neuroscience 123: 785–92.CrossRefGoogle ScholarPubMed
Fox, C. A. (1940) Certain basal telencephalic centers in the cat. Journal of Comparative Neurology 72: 1–62.CrossRefGoogle Scholar
Frank, M. G., Issa, N. P. and Stryker, M. P. (2001) Sleep enhances plasticity in the developing visual cortex. Neuron 30: 275–87.CrossRefGoogle ScholarPubMed
Freedman, J. E. and Aghajanian, G. K. (1987) Role of phosphoinositide metabolites in the prolongation of afterhyperpolarizations by alpha 1-adrenoceptors in rat dorsal raphe neurons. Journal of Neuroscience 7: 3897–906.CrossRefGoogle ScholarPubMed
Freedman, L. J., Insel, T. R. and Smith, Y. (2000) Subcortical projections of area 25 (subgenual cortex) of the macaque monkey. Journal of Comparative Neurology 29: 172–88.3.0.CO;2-8>CrossRefGoogle Scholar
Freeman, W. J. (1959) Distribution in time and space of prepyriform electrical activity. Journal of Neurophysiology 22: 644–65.CrossRefGoogle ScholarPubMed
Freeman, W. J. (1975) Mass Action in the Nervous System. New York: Academic Press.Google Scholar
Freeman, W. J. and Va, Dijk, B. W. (1988) Spatial patterns of visual cortical fast EEG during conditioned reflex in a rhesus monkey. Brain Research 422: 267–76.CrossRefGoogle Scholar
Frégnac, Y. (1999) A tale of two spikes. Nature Neuroscience 2: 299–301.CrossRefGoogle ScholarPubMed
Freund, T. F. and Buzsáki, G. (1996) Interneurons of the hippocampus. Hippocampus 6: 347–470.3.0.CO;2-I>CrossRefGoogle ScholarPubMed
Freund, T. F. and Meskenaite, V. (1992) γ-Aminobutyric acid-containing basal forebrain neurons innervate inhibitory interneurons in the neocortex. Proceedings of the National Academy of Sciences USA 89: 738–42.CrossRefGoogle ScholarPubMed
Friedman, A. and Gutnick, M. J. (1989) Intracellular calcium and control of burst generation in neurons of guinea-pig neocortex in vitro. European Journal of Neuroscience 1: 374–81.CrossRefGoogle ScholarPubMed
Fries, P., Reynolds, J. H., Rorie, A. E. and Desimone, R. (2001) Modulation of oscillatory neuronal synchronization by selective visual attention. Science 291: 1560–3.CrossRefGoogle ScholarPubMed
Fu, Y. and Shinnick-Gallagher, P. (2005) Two intra-amygdaloid pathways to the central amygdala exhibit different mechanisms of long-term potentiation. Journal of Neurophysiology 93: 3012–15.CrossRefGoogle ScholarPubMed
Fuentealba, P. and Steriade, M. (2005) The reticular nucleus revisited: intrinsic and network properties of a thalamic pacemaker. Progress in Neurobiology 75: 125–41.CrossRefGoogle ScholarPubMed
Fuentealba, P., Crochet, S., Timofeev, I. and Steriade, M. (2002) ‘Spikelets’ in cat thalamic reticular nucleus in vivo. Society for Neuroscience Abstracts 28: 144.19.Google Scholar
Fuentealba, P., Crochet, S. and Steriade, M. (2004a) The cortically evoked secondary depolarization affects the integrative properties of thalamic reticular neurons. European Journal of Neuroscience 20: 2691–6.CrossRefGoogle Scholar
Fuentealba, P., Crochet, S., Timofeev, I.et al. (2004b) Experimental evidence and modeling studies support a synchronizing role for electrical coupling in the cat thalamic reticular neurons in vivo. European Journal of Neuroscience 20: 111–19.CrossRefGoogle Scholar
Fuentealba, P., Crochet, S., Timofeev, I. and Steriade, M. (2004c) Synaptic interactions between thalamic and cortical inputs onto cortical neurons in vivo. Journal of Neurophysiology 91: 1990–8.CrossRefGoogle Scholar
Fuentealba, P., Timofeev, I. and Steriade, M. (2004d) Prolonged hyperpolarizing potentials precede spindle oscillations in the thalamic reticular nucleus. Proceedings of the National Academy of Sciences USA 101: 9816–21.CrossRefGoogle Scholar
Fuentealba, P., Timofeev, I., Bazhenov, M., Sejnowski, T. J. and Steriade, M. (2005) Membrane bistability in cat thalamic reticular neurons during spindle oscillations. Journal of Neurophysiology 93: 294–304.CrossRefGoogle Scholar
Fuster, J. M. (2004) Upper processing stages of the perception-action cycle. Trends in Cognitive Sciences 8: 143–5.CrossRefGoogle ScholarPubMed
Fuxe, K., Jacobsen, K. X., Hoistad, M.et al. (2003) The dopamine D1 receptor-rich main and paracapsular intercalated nerve cell groups of the rat amygdala: relationship to the dopamine innervation. Neuroscience 119: 733–46.CrossRefGoogle ScholarPubMed
Gaffan, D. and Murray, E. A. (1992) Monkeys (Macaca fascicularis) with rhinal cortex ablations succeed in object discrimination learning despite 24-hr intertrial intervals and fail at matching to sample despite double sample presentations. Behavioral Neuroscience 106: 30–8.CrossRefGoogle ScholarPubMed
Gais, S. and Born, J. (2004) Low acetylcholine during slow-wave sleep is critical for declarative memory consolidation. Proceedings of the National Academy of Sciences USA 101: 2140–4.CrossRefGoogle ScholarPubMed
Gais, S., Plihal, W., Wagner, U. and Born, J. (2000) Early sleep triggers memory for early visual discrimination skills. Nature Neuroscience 3: 1335–9.CrossRefGoogle ScholarPubMed
Gais, S., Mölle, M., Helms, K. and Born, J. (2002) Learning-dependent increases in sleep density. Journal of Neuroscience 22: 6830–4.CrossRefGoogle Scholar
Galarreta, M. and Hestrin, S. (1999) A network of fast-spiking cells in the neocortex connected by electrical synapses. Nature 402: 72–5.CrossRefGoogle ScholarPubMed
Galarreta, M. and Hestrin, S. (2002) Electrical and chemical synapses among parvalbumin fast-spiking GABAergic interneurons in adult mouse neocortex. Proceedings of the National Academy of Sciences USA 99: 12438–43.CrossRefGoogle ScholarPubMed
Gallagher, M., Kapp, B. S., Musty, R. E. and Driscoll, P. A. (1977) Memory formation: evidence for a specific neurochemical system in the amygdala. Science 198: 423–5.CrossRefGoogle ScholarPubMed
Garcia-Rill, E., Biedermann, J. A., Chambers, T.et al. (1995) Mesopontine neurons in schizophrenia. Neuroscience 66: 321–35.CrossRefGoogle Scholar
Gaudreau, H. and Paré, D. (1996) Projection cells of the lateral nucleus are virtually silent throughout the sleep-waking cycle. Journal of Neurophysiology 75: 1301–5.CrossRefGoogle ScholarPubMed
Gaykema, R. P., Luiten, P. G., Nyakas, C. and Traber, J. (1990) Cortical projection patterns of the medial septum-diagonal band complex. Journal of Comparative Neurology 293: 103–24.CrossRefGoogle ScholarPubMed
Gean, P. W., Huang, C. C., Lin, J. H. and Tsai, J. J. (1992) Sustained enhancement of NMDA receptor-mediated synaptic potential by isoproterenol in rat amygdalar slices. Brain Research 594: 331–4.CrossRefGoogle ScholarPubMed
Gemmell, C. and O'Mara, S. M. (2000) Long-term potentiation and paired-pulse facilitation in the prelimbic cortex of the rat following stimulation in the contralateral hemisphere in vivo. Experimental Brain Research 132: 223–9.CrossRefGoogle ScholarPubMed
Gerard, R. W. (1949) Physiology and psychiatry. American Journal of Psychiatry 106: 161–73.CrossRefGoogle ScholarPubMed
German, D. C., Manaye, K. F., Wu, D., Hersh, L. B. and Zweig, R. M. (1999) Mesopontine cholinergic and non-cholinergic neurons in schizophrenia. Neuroscience 94: 33–8.CrossRefGoogle Scholar
Gervasoni, D., Lin, S. C., Ribeiro, S.et al. (2004) Global forebrain dynamics predict rat behavioral states and their transitions. Journal of Neuroscience 24: 11137–47.CrossRefGoogle ScholarPubMed
Gewirtz, J. C., Falls, W. A. and Davis, M. (1997) Normal conditioned inhibition and extinction of freezing and fear-potentiated startle following electrolytic lesions of medical prefrontal cortex in rats. Behavioral Neuroscience 111: 712–26.CrossRefGoogle ScholarPubMed
Ghashghaei, H. T. and Barbas, H. (2002) Pathways for emotion: Interactions of prefrontal and anterior. Neuroscience 115: 1261–79.CrossRefGoogle ScholarPubMed
Gibson, J. R., Beierlein, M. and Connors, B. W. (1999) Two networks of electrically coupled inhibitory neurons in neocortex. Nature 402: 75–9.CrossRefGoogle ScholarPubMed
Gilbert, C. D. (1998) Adult cortical dynamics. Physiological Reviews 78: 467–85.CrossRefGoogle ScholarPubMed
Girod, R., Barazangi, N., McGehee, D. and Role, L. W. (2000) Facilitation of glutamatergic neurotransmission by presynaptic nicotinic acetylcholine receptors. Neuropharmacology 39: 2715–25.CrossRefGoogle ScholarPubMed
Glenn, L. L. and Steriade, M. (1982) Discharge rate and excitability of cortically projecting intralaminar thalamic neurons during waking and sleep states. Journal of Neuroscience 2: 1287–404.CrossRefGoogle ScholarPubMed
Glenn, L. L., Hada, J., Roy, J. P., Deschênes, M. and Steriade, M. (1982) Anterograde tracer and field potential analysis of the neocortical layer I projection from nucleus ventralis medialis of the thalamus in cat. Neuroscience 7: 1861–77.CrossRefGoogle ScholarPubMed
Gloor, P. (1976) Generalized and widespread bilateral paroxysmal abnormalities. In Handbook of Electroencephalography and Clinical Neurophysiology, vol. 11, Part B, ed. Remond, A., pp. 52–87. Amsterdam: Elsevier.Google Scholar
Gold, P. E. and McGaugh, J. L. (1975) A single-trace, two process view of memory storage processes. In Short-term Memory, ed. Deutsch, D. and Deutsch, J. A., pp. 355–78. New York: Academic Press.Google Scholar
Goldman, P. S. (1979) Contralateral projections to the dorsal thalamus from frontal association cortex in rhesus monkey. Brain Research 166: 166–71.CrossRefGoogle ScholarPubMed
Goldman-Rakic, P. S. (1987) Circuitry of the prefrontal cortex and the regulation of behavior by representational memory. In Handbook of Physiology, vol. 5 (The Nervous System), ed. Plum, F. and Mountcastle, V. B., pp. 373–417. Bethesda, MD: American Physiological Society.Google Scholar
Goldman-Rakic, P. S. (1988) Changing concepts of cortical connectivity: parallel distributed cortical networks. In Neurobiology of Neocortex, ed. Kakic, P. and Singer, W., pp. 177–202. New York: Wiley.Google Scholar
Golomb, D., Wang, X. J. and Rinzel, J. (1994) Synchronization properties of spindle oscillations in a thalamic reticular nucleus model. Journal of Neurophysiology 72: 1109–26.CrossRefGoogle Scholar
Golshani, P., Liu, X. B. and Jones, E. G. (2001) Differences in quantal amplitude reflect GluR4-subunit number at corticothalamic synapses on two populations of thalamic neurons. Proceedings of the National Academy of Sciences USA 98: 4172–7.CrossRefGoogle ScholarPubMed
Goosens, G. and Otto, T. A. (1998) Induction and transient suppression of long-term potentiation in the peri- and postrhinal cortices following theta-related stimulation of hippocampal field CA1. Brain Research 780: 95–101.Google Scholar
Goosens, K. A. and Maren, S. (2001) Contextual and auditory fear conditioning are mediated by the lateral, basal, and central amygdaloid nuclei in rats. Learning and Memory 8: 148–55.CrossRefGoogle ScholarPubMed
Goosens, K. A., and Maren, S. (2003) Pretraining NMDA receptor blockade in the basolateral complex. Behavioral Neuroscience 117: 738–50.CrossRefGoogle ScholarPubMed
Govindaiah and Cox, C. L. (2004) Synaptic activation of metabotropic glutamate receptors regulates dendritic outputs of thalamic interneurons. Neuron 41: 611–23.CrossRefGoogle Scholar
Grace, A. A. and Bunney, B. S. (1979) Paradoxical GABA excitation of nigral dopaminergic cells: indirect mediation through reticulata inhibitory neurons. European Journal of Pharmacology 59: 211–18.CrossRefGoogle ScholarPubMed
Grace, A. A. and Bunney, B. S. (1985) Opposing effects of striatonigral feed-back pathways on midbrain dopamine cell activity. Brain Research 333: 271–84.CrossRefGoogle Scholar
Grastyan, E., Lissak, K., Madarasz, I. and Donhoffer, H. (1959) The hippocampal electrical activity during the development of conditioned reflexes. Electroencephalography and Clinical Neurophysiology 11: 409–30.CrossRefGoogle ScholarPubMed
Gray, C. M. and McCormick, D. A. (1996) Chattering cells: superficial pyramidal neurons contributing to the generation of synchronous oscillations in the visual cortex. Science 274: 109–13.CrossRefGoogle ScholarPubMed
Gray, C. M., König, P., Engel, A. K. and Singer, W. (1989) Stimulus-specific neuronal oscillations in cat visual cortex exhibit inter-columnar synchronization which reflects global stimulus properties. Nature 338: 334–7.CrossRefGoogle Scholar
Gray, C. M., Engel, A. K., König, P. and Singer, W. (1990) Stimulus-dependent neuronal oscillations in cat visual cortex: receptive field properties and feature dependence. European Journal of Neuroscience 2: 607–19.CrossRefGoogle ScholarPubMed
Graybiel, A. M. and Ragsdale, C. W. (1979) Fiber connections of the basal ganglia. In Development and Chemical Specificity of Neurons, ed. Cuénod, M., Kreutzberg, G. W., and Bloom, F. R., pp. 239–83. Amsterdam: Elsevier.
Green, J. D. and Arduini, A. (1954) Hippocampal electrical activity in arousal. Journal of Neurophysiology 17: 533–57.CrossRefGoogle ScholarPubMed
Grenier, F., Timofeev, I. and Steriade, M. (2001) Focal synchronization of ripples (80–200 Hz) in neocortex and their neuronal correlates. Journal of Neurophysiology 86: 1884–98.CrossRefGoogle ScholarPubMed
Griffith, W. H. (1988) Membrane properties of cell types within guinea pig basal forebrain nuclei in vitro. Journal of Neurophysiology 59: 1590–612.CrossRefGoogle ScholarPubMed
Gritti, I., Mainville, L. and Jones, B. E. (1994) Projections of GABAergic and cholinergic basal forebrain and GABAergic preoptic-anterior hypothalamic neurons to the posterior lateral hypothalamus of the rat. Journal of Comparative Neurology 339: 251–68.CrossRefGoogle ScholarPubMed
Gross, D. W. and Gotman, J. (1999) Correlation of high-frequency oscillations with the sleep-wake cycle and cognitive activity in humans. Neuroscience 94: 1005–18.CrossRefGoogle ScholarPubMed
Guido, W. and Weyand, T. (1995) Burst responses in thalamic relay cells of the awake, behaving cat. Journal of Neurophysiology 74: 1782–6.CrossRefGoogle ScholarPubMed
Guido, W., Lu, S. M. and Sherman, S. M. (1992) Relative contributions of burst and tonic responses to the receptive field properties of lateral geniculate neurons in the cat. Journal of Neurophysiology 68: 2199–211.CrossRefGoogle ScholarPubMed
Guillery, R. W. (1969) The organization of synaptic interconnections in the laminae of the dorsal lateral geniculate nucleus of the cat. Zeitschrift für Zellforschung und Mikroskopische Anatomie 96: 1–38.CrossRefGoogle ScholarPubMed
Guillery, R. W. and Harting, J. K. (2002) Structure and connections of the thalamic reticular nucleus: advancing views over half of century. Journal of Comparative Neurology 463: 360–71.CrossRefGoogle Scholar
Gulyás, A. I., Hájos, N. and Freund, T. F. (1996) Interneurons containing calretinin are specialized to control other interneurons in the rat hippocampus. Journal of Neuroscience 16: 3397–411.CrossRefGoogle ScholarPubMed
Gupta, A., Wang, Y. and Markram, H. (2000) Organizing principles for a diversity of GABAergic interneurons and synapses in the neocortex. Science 287: 273–8.CrossRefGoogle ScholarPubMed
Gurden, H., Tassin, J. P. and Jay, T. M. (1999) Integrity of the mesocortical dopaminergic system is necessary for complete expression of in vivo hippocampal-prefrontal cortex long-term potentiation. Neuroscience 94: 1019–27.CrossRefGoogle ScholarPubMed
Gutnick, M. J., Heinemann, U. and Prince, D. A. (1979) Stimulus induced and seizure related changes in extracellular potassium concentration in cat thalamus (VPL). Electroencephalography and Clinical Neurophysiology 47: 329–44.CrossRefGoogle Scholar
Halgren, E., Bab, T. L. and Crandall, P. H. (1978) Human hippocampal formation EEG desynchronizes during attentiveness and movement. Electroencephalography and Clinical Neurophysiology 44: 778–81.CrossRefGoogle ScholarPubMed
Halgren, E., Boujon, C., Clarke, J., Wang, C. and Chauvel, P. (2002) Rapid distributed fronto-parieto-occipital processing stages during working memory in humans. Cerebral Cortex 12: 710–28.CrossRefGoogle ScholarPubMed
Hall, E. (1972a) The amygdala of the cat: A Golgi study. Zeitschrift für Zellforschung 134: 439–58.CrossRefGoogle Scholar
Hall, E. (1972b) Some aspects of the structural organization of the amygdala. In The Neurobiology of the Amygdala, ed. Eleftheriou, B. E., pp. 95–121. New York: Plenum Press.CrossRefGoogle Scholar
Hall, G. B. C., Szechtman, H. and Nahmias, C. (2003) Enhanced salience and emotion recognition in autism: a PET study. American Journal of Psychiatry 160: 1439–41.CrossRefGoogle ScholarPubMed
Halliwell, J. V. (1986) M-current in human neocortical neurones. Neuroscience Letters 67: 1–6.CrossRefGoogle ScholarPubMed
Hamam, B. N., Kennedy, T. E., Alonso, A. and Amaral, D. G. (2000) Morphological and electrophysiological characteristics of layer V neurons of the rat medial entorhinal cortex. Journal of Comparative Neurology 418: 457–72.3.0.CO;2-L>CrossRefGoogle Scholar
Hamann, S. B., Ely, T. D., Grafton, S. T. and Kilts, C. D. (1999) Amygdala activity related to enhanced memory for pleasant and aversive stimuli. Nature Neuroscience 2: 289–93.CrossRefGoogle ScholarPubMed
Hammond, C. and Crépel, F. (1992) Evidence for a slowly inactivating K+ current in prefrontal cortical cells. European Journal of Neuroscience 4: 1087–92.CrossRefGoogle ScholarPubMed
Hardy, S. G. and Leichnetz, G. R. (1981) Frontal cortical projections to the periaqueductal gray in the rat: a retrograde and orthograde horseradish peroxidase study. Neuroscience Letters 23: 13–17.CrossRefGoogle ScholarPubMed
Hartveit, E. and Heggelund, P. (1994) Response variability of single cells in the dorsal lateral geniculate nucleus of the cat. Comparison with retinal input and effect of brain stem stimulation. Journal of Neurophysiology 72: 1278–89.CrossRefGoogle ScholarPubMed
Hashikawa, T., Rauseell, E., Molinari, M. and Jones, E. G. (1991) Parvalbumin- and calbindin-containing neurons in the monkey medial geniculate complex: differential distribution and cortical layer specific projections. Brain Research 544: 335–41.CrossRefGoogle ScholarPubMed
Hasselmo, M. E. (1999) Neuromodulation: acetylcholine and memory consolidation. Trends in Cognitive Sciences 3: 351–9.CrossRefGoogle ScholarPubMed
Hasue, R. H. and Shammah-Lagnado, S. J. (2002) Origin of the dopaminergic innervation of the central extended amygdala and accumbens shell: a combined retrograde tracing and immunohistochemical study in the rat. Journal of Comparative Neurology 454: 15–33.CrossRefGoogle ScholarPubMed
Hatfield, T. and McGaugh, J. L. (1999) Norepinephrine infused into the basolateral amygdala posttraining enhances retention in a spatial water maze task. Neurobiology of Learning and Memory 71: 232–9.CrossRefGoogle Scholar
He, J. (2003) Slow oscillation in non-lemniscal auditory thalamus. Journal of Neuroscience 23: 8281–90.CrossRefGoogle ScholarPubMed
Herbert, M. R., Ziegler, D. A., Deutsch, C. K.et al. (2003) Dissociations of cerebral cortex, subcortical and cerebral white matter volumes in autistic boys. Brain 126: 1182–92.CrossRefGoogle ScholarPubMed
Herculano-Houzel, S., Munk, M. H. J., Neuenschwander, S. and Singer, W. (1999) Precisely synchronized oscillatory firing patterns require electroencephalographic activation. Journal of Neuroscience 19: 3992–4010.CrossRefGoogle ScholarPubMed
Herkenham, M. (1978) The connections of the nucleus reuniens thalami: evidence for a direct thalamo-hippocampal pathway in the rat. Journal of Comparative Neurology 177: 589–610.CrossRefGoogle ScholarPubMed
Herkenham, M. (1979) The afferent and efferent connections of the ventromedial thalamic nucleus in the rat. Journal of Comparative Neurology 183: 487–518.CrossRefGoogle ScholarPubMed
Hernández-Cruz, A. and Pape, H. C. (1989) Identification of two calcium currents in acutely dissociated neurons from the rat lateral geniculate nucleus. Journal of Neurophysiology 61: 1270–83.CrossRefGoogle ScholarPubMed
Herry, C. and Garcia, R. (2003) Behavioral and paired-pulse facilitation analyses of long-lasting depression at excitatory synapses in the medial prefrontal cortex in mice. Behavioral Brain Research 146: 89–96.CrossRefGoogle ScholarPubMed
Hestrin, S. and Armstrong, W. E. (1996) Morphology and physiology of cortical neurons in layer I. Journal of Neuroscience 16: 5290–300.CrossRefGoogle ScholarPubMed
Hicks, T. P., Metherate, R., Landry, P. and Dykes, R. W. (1986) Bicuculline-induced alterations of response properties in functionally identified ventroposteiror thalamic neurones. Experimental Brain Research 63: 248–64.CrossRefGoogle ScholarPubMed
Hikosaka, O. and Wurtz, R. H. (1983) Visual and oculomotor function of monkey substantia nigra pars reticulata, IV: relation of substantia nigra to superior colliculus. Journal of Neurophysiology 49: 1285–301.CrossRefGoogle ScholarPubMed
Hirsch, J. C. and Burnod, Y. (1987) A synaptically evoked late hyperpolarization in the rat dorsolateral geniculate neurons in vitro. Neuroscience 23: 457–68.CrossRefGoogle ScholarPubMed
Hirsch, J. C., Fourment, A. and Marc, M. E. (1983) Sleep-related variations of membrane potential in the lateral geniculate body relay neurons of the cat. Brain Research 259: 308–12.CrossRefGoogle ScholarPubMed
, N. and Destexhe, A. (2000) Synaptic background activity enhances the responsiveness of neocortical pyramidal neurons. Journal of Neurophysiology 84: 1488–96.CrossRefGoogle ScholarPubMed
Hobin, J. A., Goosens, K. A. and Maren, S. (2003) Context-dependent neuronal activity in the lateral amygdala represents fear memories after extinction. Journal of Neuroscience 23: 8410–16.CrossRefGoogle ScholarPubMed
Hobson, J. A. (1988) The Dreaming Brain. New York: Basic Books.Google Scholar
Hobson, J. A. and Pace-Schott, E. F. (2002) The cognitive neuroscience of sleep: neuronal systems, consciousness and learning. Nature Reviews Neuroscience 3: 679–93.CrossRefGoogle Scholar
Hobson, J. A. and Steriade, M. (1986) Neuronal basis of behavioral state control. In Handbook of Physiology, ed. Mountcastle, V. B. and Bloom, F. E., section 1, vol. IV, pp. 701–823. Bethesda, MD: American Physiological Society.Google Scholar
Hobson, J. A., Pace-Schott, E. and Stickgold, R. (2000) Dreaming and the brain: toward a cognitive neuroscience of conscious states. Brain and Behavioral Sciences 23: 793–842.CrossRefGoogle Scholar
Hofle, N., Paus, T., Reutens, D.et al. (1997) Regional cerebral blood flow changes as a function of delta and spindle activity during slow wave sleep in humans. Journal of Neuroscience 17: 4800–8.CrossRefGoogle ScholarPubMed
Holahan, M. R. and White, N. M. (2002) Conditioned memory modulation, freezing, and avoidance as measures of amygdala-mediated conditioned fear. Neurobiology of Learning and Memory 77: 250–75.CrossRefGoogle ScholarPubMed
Holstege, G. (1990) Subcortical limbic system projections to caudal brainstem and spinal cord. In The Human Nervous System, ed. Paxinos, G., pp. 261–86. New York: Academic Press.Google Scholar
Holstege, G., Bandler, R. and Saper, C. B. (1996) The emotional motor system. Progress in Brain Research 107: 3–6.CrossRefGoogle ScholarPubMed
Homma, Y., Skinner, R. D. and Garcia-Rill, E. (2002) Effects of pedunculopontine nucleus (PPN) stimulation on caudal pontine reticular formation (PnC) neurons in vitro. Journal of Neurophysiology 87: 3033–47.CrossRefGoogle ScholarPubMed
Honda, T. and Semba, K. (1995) An ultrastructural study of cholinergic and non-cholinergic neurons in the laterodorsal and pedunculopontine tegmental nuclei in the rat. Neuroscience 68: 837–53.CrossRefGoogle ScholarPubMed
Hope, B., Michael, G., Knigge, K. and Vincent, S. (1991) Neuronal NADPH diaphorase is a nitric oxide synthase. Proceedings of the National Academy of Sciences USA 88: 2811–14.CrossRefGoogle ScholarPubMed
Hopkins, D. A. and Holstege, G. (1978) Amygdaloid projections to the mesencephalon, pons and medulla oblongata in the cat. Experimental Brain Research 32: 529–47.CrossRefGoogle ScholarPubMed
Hou, Y. P., Manns, I. D. and Jones, B. E. (2002) Immunostaining of cholinergic pontomesencephalic neurons for α1 versus α2 adrenergic receptors suggests different sleep-wake state activities and roles. Neuroscience 114: 517–21.CrossRefGoogle Scholar
Hu, B. (1995) Cellular basis of temporal synaptic signaling: an in vitro electrophysiological study in rat auditory thalamus. Journal of Physiology (London) 483: 167–82.CrossRefGoogle Scholar
Hu, B., Bouhassira, D., Steriade, M. and Deschênes, M. (1988) The blockage of ponto-geniculo-occipital waves in the cat lateral geniculate nucleus by nicotinic antagonists. Brain Research 473: 394–7.CrossRefGoogle ScholarPubMed
Hu, B., Steriade, M. and Deschênes, M. (1989a) The effects of peribrachial stimulation on reticular thalamic neurons: the blockage of spindle waves. Neuroscience 31: 1–12.CrossRefGoogle Scholar
Hu, B., Steriade, M. and Deschênes, M. (1989b) The effects of brainstem peribrachial stimulation on neurons of the lateral geniculate nucleus. Neuroscience 31: 13–24.CrossRefGoogle Scholar
Hu, B., Steriade, M. and Deschênes, M. (1989c) The cellular mechanism of thalamic ponto-geniculo-occipital waves. Neuroscience 31: 25–35.CrossRefGoogle Scholar
Huang, C. C., Hsu, K. S. and Gean, P. W. (1996) Isoproterenol potentiates synaptic transmission primarily by enhancing presynaptic calcium influx via P- and/or Q-type calcium channels in the rat amygdala. Journal of Neuroscience 16: 1026–33.CrossRefGoogle ScholarPubMed
Huang, Y. Y. and Kandel, E. R. (1998) Postsynaptic induction and PKA-dependent expression of LTP in the lateral amygdala. Neuron 21: 169–78.CrossRefGoogle ScholarPubMed
Hubel, D. H. and Wiesel, T. N. (1961) Integrative action in the cat's lateral geniculate body. Journal of Physiology (London) 155: 385–98.CrossRefGoogle ScholarPubMed
Huber, R., Ghilardi, M. F., Massimini, M. and Tononi, G. (2004) Local sleep and learning. Nature 430: 78–81.CrossRefGoogle Scholar
Hugues, S., Deschaux, O. and Garcia, R. (2004) Postextinction infusion of a mitogen-activated protein kinase inhibitor into the medial prefrontal cortex impairs memory of the extinction of conditioned fear. Learning and Memory 11: 540–3.CrossRefGoogle ScholarPubMed
Hughes, S. W., Cope, D. W., Toth, T. I., Williams, S. R. and Crunelli, V. (1999) All thalamocortical neurones possess a T-type Ca2+ ‘window’ current that enables the expression of bistability-mediated activities. Journal of Physiology (London) 517: 805–15.CrossRefGoogle ScholarPubMed
Hughes, S. W., Blethyn, K. L., Cope, D. W. and Crunelli, V. (2002a) Properties and origin of spikelets in thalamocortical neurones in vitro. Neuroscience 110: 395–401.CrossRefGoogle Scholar
Hughes, S. W., Cope, D. W., Blethyn, K. L. and Crunelli, V. (2002b) Cellular mechanisms of the slow (<1 Hz) oscillation in thalamocortical neurons in vitro. Neuron 33: 947–58.CrossRefGoogle Scholar
Hughes, S. W., Lörincz, M., Cope, D. W.et al. (2004) Synchronized oscillations at alpha and theta frequencies in the lateral geniculate nucleus. Neuron 42: 253–68.CrossRefGoogle ScholarPubMed
Huguenard, J. R. (1996) Low-threshold calcium currents in central nervous system neurons. Annual Review of Physiology 58: 329–48.CrossRefGoogle ScholarPubMed
Huguenard, J. R. and Prince, D. A. (1992) A novel T-type current underlies prolonged Ca2+-dependent burst firing in GABAergic neurons of rat thalamic reticular nucleus. Journal of Neuroscience 12: 3804–17.CrossRefGoogle Scholar
Huguenard, J. R. and Prince, D. A. (1994) Intrathalamic rhythmicity studied in vitro: nominal T current modulation causes robust anti-oscillatory effects. Journal of Neuroscience 14: 5485–502.CrossRefGoogle Scholar
Hugues, S., Deschaux, O. and Garcia, R. (2004) Postextinction infusion of a mitogen-activated protein kinase inhibitor into the medial prefrontal cortex impairs memory of the extinction of conditioned fear. Learning and Memory 11: 540–3.CrossRefGoogle ScholarPubMed
Humeau, Y., Shaban, H., Bissière, S. and Lüthi, A. (2003) Presynaptic induction of heterosynaptic associative plasticity in the mammalian brain. Nature 426: 841–5.CrossRefGoogle ScholarPubMed
Humeau, Y., Herry, C., Kemp, N.et al. (2005) Dendritic spine heterogeneity determines afferent-specific Hebbian plasticity in the amygdala. Neuron 45: 119–31.CrossRefGoogle ScholarPubMed
Humphrey, T. (1936) The telencephalon of the bat. I. The noncortical nuclear masses and certain pertinent fiber connections. Journal of Comparative Neurology 65: 603–711.CrossRefGoogle Scholar
Hunt, C. A., Pang, D. Z. and Jones, E. G. (1991) Distribution and density of GABA cells in intralaminar and adjacent nuclei of monkey thalamus. Neuroscience 43: 185–96.CrossRefGoogle ScholarPubMed
Hurley, K. M., Herbert, H., Moga, M. M. and Saper, C. B. (1991) Efferent projections of the infralimbic cortex of the rat. Journal of Comparative Neurology 308: 249–76.CrossRefGoogle ScholarPubMed
Hutcheon, B., Miura, R. M. and Puil, E. (1996) Subthreshold membrane resonance in neocortical neurons. Journal of Neurophysiology 76: 683–97.CrossRefGoogle ScholarPubMed
Inglis, W. L. and Semba, K. (1996) Colocalization of ionotropic glutamate receptor subunits with NADPH-diaphorase-containing neurons in the rat mesopontine tegmentum. Journal of Comparative Neurology 368: 17–32.3.0.CO;2-N>CrossRefGoogle ScholarPubMed
Ingvar, D. H., Sjölund, B. and Ardo, A. (1976) Correlation between ECG frequency, cerebral oxygen uptake and blood flow. Electroencephalography and Clinical Neurophysiology 41: 268–76.CrossRefGoogle Scholar
Insausti, R., Amaral, D. G. and Cowan, W. M. (1987) The entorhinal cortex of the monkey: III. Subcortical afferents. Journal of Comparative Neurology 264: 396–408.CrossRefGoogle ScholarPubMed
Introini-Collison, I. B., Miyazaki, B. and McGaugh, J. L. (1991) Involvement of the amygdala in the memory-enhancing effects of clenbuterol. Psychopharmacology 104: 541–4.CrossRefGoogle ScholarPubMed
Inubushi, S., Kobayashi, T., Oshima, T. and Torii, S. (1978a) Intracellular recordings from motor cortex during EEG arousal in unanesthetized brain preparations of the cat. Japanese Journal of Physiology 28: 669–88.CrossRefGoogle Scholar
Inubushi, S., Kobayashi, T., Oshima, T. and Torii, S. (1978b) An intracellular analysis of EEG arousal in cat motor cortex. Japanese Journal of Physiology 28: 689–708.CrossRefGoogle Scholar
Irle, E. and Markowitsch, H. J. (1984) Basal forebrain efferents reach the whole cerebral cortex of the cat. Brain Research Bulletin 12: 493–512.CrossRefGoogle ScholarPubMed
Ito, K. and McCarley, R. W. (1984) Alterations in membrane potential and excitability of cat medial pontine reticular formation neurons during changes in naturally occurring sleep-wake states. Brain Research 292: 169–75.CrossRefGoogle ScholarPubMed
Iwata, J., LeDoux, J. E., Meeley, M. P., Arneric, S. and Reis, D. J. (1986) Intrinsic neurons in the amygdaloid field projected to by the medial geniculate body mediate emotional responses conditioned to acoustic stimuli. Brain Research 383: 195–214.CrossRefGoogle ScholarPubMed
Izquierdo, I. and Medina, J. H. (1993) Role of the amygdala, hippocampus and entorhinal cortex in memory consolidation and expression. Brazilian Journal of Medical and Biological Research 26: 573–89.Google ScholarPubMed
Jacobs, B. L. and McGinty, D. J. (1971) Amygdala unit activity during sleep and waking. Experimental Neurology 33: 1–15.CrossRefGoogle ScholarPubMed
Jahnsen, H. and Llinás, R. (1984a) Electrophysiological properties of guinea-pig thalamic neurones: an in vitro study. Journal of Physiology (London) 349: 205–26.CrossRefGoogle Scholar
Jahnsen, H. and Llinás, R. (1984b) Ionic basis for electroresponsiveness and oscillatory properties of guinea-pig thalamic neurones in vitro. Journal of Physiology (London) 349: 227–47.CrossRefGoogle Scholar
Jasper, H. H. (1949) Diffuse projection systems: the integrative action of the thalamic reticular system. Electroencephalography and Clinical Neurophysiology 1: 405–20.CrossRefGoogle ScholarPubMed
Jasper, H. H. (1958) Recent advances in our understanding of ascending activities of the reticular system. In Reticular Formation of the Brain, ed. Jasper, H. H., Proctor, L. D., Knighton, R. S., Noshay, W. C. and Costello, R. T., pp. 319–31. Boston, Toronto: Little-Brown.Google Scholar
Jay, T. M., Burette, F. and Laroche, S. (1996) Plasticity of the hippocampal-prefrontal cortex synapses. Journal of Physiology (Paris) 90: 361–6.CrossRefGoogle ScholarPubMed
Jeanmonod, D., Magnin, M. and Morel, A. (1996) Low-threshold calcium spike-bursts in the human thalamus: common physiopathology for sensory, motor and limbic positive symptoms. Brain 119: 363–75.CrossRefGoogle ScholarPubMed
Jenkins, J. and Dallenback, K. (1924) Oblivescence during sleep and waking. American Journal of Psychology 35: 605–12.CrossRefGoogle Scholar
Johnston, J. B. (1923) Further contributions to the study of the evolution of the forebrain. Journal of Comparative Neurology 35: 337–481.CrossRefGoogle Scholar
Jolkkonen, E. and Pitkänen, A. (1998) Intrinsic connections of the rat amygdaloid complex: projections originating in the central nucleus. Journal of Comparative Neurology 395: 53–72.3.0.CO;2-G>CrossRefGoogle ScholarPubMed
Jolkkonen, E., Miettinen, R., Pikkarainen, M. and Pitkänen, A. (2002) Projections from the amygdaloid complex to the magnocellular cholinergic basal forebrain in rat. Neuroscience 111: 133–49.CrossRefGoogle ScholarPubMed
Jones, B. E. (2003) Arousal systems. Frontiers in Bioscience 8: s438–51.CrossRefGoogle ScholarPubMed
Jones, B. E. (2004) Activity, modulation and role of basal forebrain cholinergic neurons innervating the cerebral cortex. In Acetylcholine in the Cerebral Cortex (Progress in Brain Research, vol. 145), ed. Descarries, L., Krnjević, K. and Steriade, M., pp. 157–69. Amsterdam: Elsevier.Google Scholar
Jones, B. E. and Beaudet, A. (1987) Distribution of acetylcholine and catecholamine neurons in the cat brain stem studied by choline acetyltransferase and tyrosine hydroxylase immunohistochemistry. Journal of Comparative Neurology 261: 15–32.CrossRefGoogle Scholar
Jones, B. E., and Cuello, A. C. (1989) Afferents to the basal forebrain cholinergic cell area from pontomesencephalic — catecholamine, serotonin, and acetylcholine — neurons. Neuroscience 31: 37–61.CrossRefGoogle ScholarPubMed
Jones, B. E. and Moore, R. Y. (1977) Ascending projections of the locus coeruleus in the rat. II. Autoradiographic study. Brain Research 127: 25–53.CrossRefGoogle ScholarPubMed
Jones, E. G. (1975) Varieties and distribution of non-pyramidal cells in the somatic sensory cortex of the squirrel monkey. Journal of Comparative Neurology 160: 205–68.CrossRefGoogle ScholarPubMed
Jones, E. G. (1984) Laminar distribution of cortical efferent cells. In Cerebral Cortex, vol. 1 (Cellular Components of the Cerebral Cortex), ed. Peters, A. and Jones, E. G., pp. 521–33. New York: Plenum.Google Scholar
Jones, E. G. (1985) The Thalamus. New York: Plenum.CrossRefGoogle Scholar
Jones, E. G. (1991) Cellular organization in the primate postcentral gyrus. In Information Processing in the Somatosensory System, ed. Franzen, O. and Westman, J., pp. 95–107. New York: Macmillan.CrossRefGoogle Scholar
Jones, E. G. (1997). A description of the human thalamus. In Thalamus, vol. 2 (Experimental and Clinical Aspects), ed. Steriade, M., Jones, E. G. and McCormick, D. A., pp. 425–99. Amsterdam: Elsevier.Google Scholar
Jones, E. G. (2000) Cortical and subcortical contributions to activity-dependent plasticity in primate somatosensory cortex. Annual Reviews of Neuroscience 23: 1–37.CrossRefGoogle ScholarPubMed
Jones, E. G. (2002) Thalamic circuitry and thalamocortical synchrony. Philosophical Transactions of the Royal Society of LondonB357: 1659–73.CrossRefGoogle ScholarPubMed
Jones, E. G. and Pons, T. P. (1998) Thalamic and brainstem contributions to large-scale plasticity of primate somatosensory cortex. Science 282: 1121–25.CrossRefGoogle ScholarPubMed
Jones, E. G. and Powell, T. P. S. (1970) An anatomical study of converging sensory pathways within the cerebral cortex of the monkey. Brain 93: 793–820.CrossRefGoogle ScholarPubMed
Jones, R. S. G. (1990) Intrinsic properties of neurones in layer II of the rat entorhinal cortex in vitro. Journal of Physiology (London) 425: 86P.Google Scholar
Jones, R. S. G. (1994) Synaptic and intrinsic properties of neurons of origin of the perforant path in layer II of the rat entorhinal cortex in vitro. Hippocampus 4: 335–53.CrossRefGoogle ScholarPubMed
Jones, R. S. G. and Bühl, E. H. (1993) Basket-like interneurones in layer II of the entorhinal cortex exhibit a powerful NMDA-mediated synaptic excitation. Neuroscience Letters 149: 35–9.CrossRefGoogle ScholarPubMed
Jongen-Rêlo, A. L. and Amaral, D. G. (1998) Evidence for a GABAergic projection from the central nucleus of the amygdala to the brainstem of the macaque monkey: a combined retrograde tracing and in situ hybridization study. European Journal of Neuroscience 10: 2924–33.CrossRefGoogle ScholarPubMed
Jouvet, M. (1965) Paradoxical sleep – a study of its nature and mechanisms. Progress in Brain Research 18: 20–57.CrossRefGoogle ScholarPubMed
Jouvet, M. (1972) The role of monoamines and acetylcholine-containing neurons in the regulation of the sleep-waking cycle. Ergebnisse der Physiologie 64: 166–307.Google ScholarPubMed
Jouvet, M. and Delorme, J. F. (1965) Locus coeruleus et sommeil paradoxal. Comptes Rendus de la Société de Biologie de Paris 159: 895–9.Google Scholar
Jouvet, M. and Michel, F. (1959) Corrélations electromyographiques du sommeil chez le chat décortiqué et mésencephalique chronique. Comptes Rendus de la Société de Biologie de Paris 153: 422–5.Google Scholar
Kamal, A. M. and Tömböl, T. (1975) Golgi studies on the amygdaloid nuclei of the cat. Journal für Hirnforschung 16: 175–201.Google ScholarPubMed
Kammermeier, P. J. and Jones, S. W. (1997) High-voltage-activated calcium currents in neurons acutely isolated from the ventrobasal nucleus of the rat thalamus. Journal of Neurophysiology 77: 465–75.CrossRefGoogle ScholarPubMed
Kamondi, A., Williams, J. A., Hutcheon, B. and Reiner, P. B. (1992) Membrane properties of mesopontine cholinergic neurons studied with the whole-cell patch-clamp technique: implications for behavioral state control. Journal of Neurophysiology 68: 1359–72.CrossRefGoogle ScholarPubMed
Kang, Y. and Kayano, F. (1994) Electrophysiological and morphological characteristics of layer VI pyramidal cells in the cat motor cortex. Journal of Neurophysiology 72: 578–91.CrossRefGoogle ScholarPubMed
Kang, Y. and Kitai, S. T. (1990) Electrophysiological properties of pedunculopontine neurons and their postsynaptic responses following stimulation of substantia nigra reticulata. Brain Research 535: 79–95.CrossRefGoogle ScholarPubMed
Kapp, B. S., Frysinger, R. C., Gallagher, M. and Haselton, J. R. (1979) Amygdala central nucleus lesions: Effects on heart rate conditioning in the rabbit. Physiology and Behavior 23: 1109–17.CrossRefGoogle Scholar
Kapp, B. S., Supple, W. J. and Whalen, P. J. (1994) Effects of electrical stimulation of the amygdaloid central nucleus on neocortical arousal in the rabbit. Behavioral Neuroscience 108: 81–93.CrossRefGoogle ScholarPubMed
Kasamatsu, T. (1976) Visual cortical neurons influenced by the oculomotor input: characterization of their receptive field properties. Brain Research 113: 271–92.CrossRefGoogle ScholarPubMed
Kasper, E. M., Larkman, A. U., Lübke, J. and Blakemore, C. (1994a) Pyramidal neurons in layer 5 of the rat visual cortex. I. Correlation among cell morphology, intrinsic electrophysiological properties, and axon targets. Journal of Comparative Neurology 339: 459–74.CrossRefGoogle Scholar
Kasper, E. M., Lübke, J., Larkman, A. U. and Blakemore, C. (1994b) Pyramidal neurons in layer 5 of the rat visual cortex. III. Differential maturation of axon targeting, dendritic morphology, and electrophysiological properties. Journal of Comparative Neurology 339: 495–518.CrossRefGoogle Scholar
Kato, N. (1990) Cortico-thalamo-cortical projection between visual cortices. Brain Research 509: 150–2.CrossRefGoogle ScholarPubMed
Kawaguchi, Y. (1993) Groupings of nonpyramidal and pyramidal cells with specific physiological and morphological characteristics in rat frontal cortex. Journal of Neurophysiology 69: 416–31.CrossRefGoogle ScholarPubMed
Kawaguchi, Y. (2001) Distinct firing patterns of neuronal subtypes in cortical synchronized activities. Journal of Neuroscience 21: 7261–72.CrossRefGoogle ScholarPubMed
Kawaguchi, Y. and Kubota, Y. (1993) Correlation of physiological subgroupings of nonpyramidal cells with parvalbumine- and calbindinD28k-immunoreactive neurons in layer V of rat frontal cortex. Journal of Neurophysiology 70: 387–96.CrossRefGoogle Scholar
Kawaguchi, Y. and Kubota, Y. (1997) GABAergic cell subtypes and their synaptic connections in rat frontal cortex. Cerebral Cortex 7: 476–86.CrossRefGoogle ScholarPubMed
Kayama, Y., Ohta, M. and Jodo, E. (1992) Firing of ‘possibly’ cholinergic neurons in the rat laterodorsal tegmental nucleus during sleep and wakefulness. Brain Research 569: 210–20.CrossRefGoogle ScholarPubMed
Kemp, N. and Bashir, Z. I. (2001) Long-term depression: a cascade of induction and expression mechanisms. Progress in Neurobiology 65: 339–65.CrossRefGoogle ScholarPubMed
Kemppainen, S. and Pitkänen, A. (2000) Distribution of parvalbumin, calretinin, and calbindin-D28k immunoreactivity in the rat amygdaloid complex and colocalization with gamma-aminobutyric acid. Journal of Comparative Neurology 426: 441–67.3.0.CO;2-7>CrossRefGoogle Scholar
Kettenmann, H. and Schachner, M. (1985) Pharmacological properties of γ-aminobutyric acid-, glutamate-, and aspartate-induced depolarizations in cultured astrocytes, Journal of Neuroscience 5: 3295–301.CrossRefGoogle ScholarPubMed
Khateb, A., Mühlethaler, M., Alonso, A.et al. (1992) Cholinergic nucleus basalis neurons display the capacity for rhythmic bursting activity mediated by low-threshold calcium spikes. Neuroscience 51: 489–94.CrossRefGoogle ScholarPubMed
Khateb, A., Fort, P., Serafin, M., Jones, B. E. and Mühlethaler, M. (1995a) Rhythmical bursts induced by NMDA in guinea-pig cholinergic nucleus basalis neurones in vitro. Journal of Physiology (London) 487: 623–38.CrossRefGoogle Scholar
Khateb, A., Fort, P., Pegna, A., Jones, B. E. and Mühlethaler, M. (1995b) Cholinergic nucleus basalis neurons are excited by histamine in vitro. Neuroscience 69: 495–506.CrossRefGoogle Scholar
Khateb, A., Fort, P., Williams, S.et al. (1997) Modulation of cholinergic nucleus basalis neurons by acetylcholine and N-methyl-D-aspartate. Neuroscience 81: 47–55.CrossRefGoogle ScholarPubMed
Killcross, S., Robbins, T. W. and Everitt, B. J. (1997) Different types of fear-conditioned behaviour mediated by separate nuclei within amygdala. Nature 388: 377–80.CrossRefGoogle ScholarPubMed
Kim, J. J. and Fanselow, M. S. (1992) Modality-specific retrograde amnesia of fear. Science 256: 675–7.CrossRefGoogle Scholar
Kim, M. J., Chun, S. K., Kim, Y. B., Mook-Jung, I. and Jung, M. W. (2003) Long-term potentiation in visual cortical projections to the medial prefrontal cortex of the rat. Neuroscience 120: 283–9.CrossRefGoogle ScholarPubMed
Kim, U. and McCormick, D. A. (1998) Functional and ionic properties of a slow afterhyperpolarization in ferret perigeniculate neurones in vitro. Journal of Neurophysiology 80: 1222–35.CrossRefGoogle Scholar
Kim, U., Bal, T. and McCormick, D. A. (1995) Spindle waves are propagating synchronized oscillations in the ferret LGNd in vitro. Journal of Neurophysiology 74: 1301–23.CrossRefGoogle ScholarPubMed
Kinney, H. C., Korein, J., Panigrahy, A., Dikkes, P. and Goode, R. (1994) Neuropathological findings in the brain of Karen Ann Quinlan. New England Journal of Medicine 330: 1469–75.CrossRefGoogle ScholarPubMed
Kinomura, S., Larsson, J., Gulyás, B. and Roland, P. (1996) Activation by attention of the human reticular formation and thalamic intralaminar nuclei. Science 271: 512–15.CrossRefGoogle ScholarPubMed
Kisvárday, Z. F., Beaulieu, C. and Eysel, U. T. (1993) Network of GABAergic large basket cells in cat visual cortex (area 18): implication for lateral disinhibition. Journal of Comparative Neurology 327: 398–415.CrossRefGoogle ScholarPubMed
Kitsikis, A. and Steriade, M. (1981) Immediate behavioral effects of kainic acid injections into the midbrain reticular core. Behavioral Brain Research 3: 361–80.CrossRefGoogle ScholarPubMed
Kleitman, N. (1963) Sleep and Wakefulness. Chicago: University of Chicago Press.Google Scholar
Klink, R. and Alonso, A. (1993) Ionic mechanisms for the subthreshold oscillations and differential electroresponsiveness of medial entorhinal cortex layer II neurons. Journal of Neurophysiology 70: 149–57.CrossRefGoogle ScholarPubMed
Kobayashi, Y., Inoue, Y., Yamamoto, M., Isa, T. and Aizawa, H. (2001) Contribution of pedunculopontine tegmental nucleus neurons to performance of visually guided saccade tasks in monkeys. Journal of Neurophysiology 88: 715–31.CrossRefGoogle Scholar
Koch, C. (1998) The neuroanatomy of visual consciousness. In Consciousness: At the Frontiers of Neuroscience (Advances in Neurology, vol. 77), ed. Jasper, H. H., Descarries, L., Castelucci, V. F. and Rossignol, S., pp. 229–41. Philadelphia, PA: Lippincott-Raven.Google Scholar
Köhler, C. and Steinbusch, H. (1982) Identification of serotonin and non-serotonin-containing neurons of the mid-brain raphe projecting to the entorhinal area and the hippocampal formation. A combined immunohistochemical and fluorescent retrograde tracing study in the rat brain. Neuroscience 7: 951–75.CrossRefGoogle ScholarPubMed
Köhler, C., Chan-Palay, V., Haglund, L. and Steinbusch, H. (1980) Immunohistochemical localization of serotonin nerve terminals in the lateral entorhinal cortex of the rat: demonstration of two separate patterns of innervation from the midbrain raphe. Anatomy and Embryology 160: 121–9.CrossRefGoogle ScholarPubMed
Köhler, C., Chan-Palay, V. and Steinbusch, H. (1981) The distribution and orientation of serotonin fibers in the entorhinal and other retrohippocampal areas. An immunohistochemical study with anti-serotonin antibodies in the rat brain. Anatomy and Embryology 161: 237–64.CrossRefGoogle Scholar
Köhler, C., Swanson, L. W., Haglund, L. and Wu, J. Y. (1985a) The cytoarchitecture, histochemistry and projections of the tuberomammillary nucleus in the rat. Neuroscience 16: 85–110.CrossRefGoogle Scholar
Köhler, C., Wu, J. Y. and Chan-Palay, V. (1985b) Neurons and terminals in the retrohippocampal region in he rat's brain identified by anti-gamma-aminobutyric acid and anti-glutamic acid decarboxylase immunocytochemistry. Anatomy and Embryology (Berlin) 173: 35–44.CrossRefGoogle Scholar
Kohlmeier, K. A. and Reiner, P. B. (1999) Noradrenaline excites non-cholinergic laterodorsal tegmental neurons via two distinct mechanisms. Neuroscience 93: 619–30.CrossRefGoogle ScholarPubMed
Koo, J. W., Han, J. S. and Kim, J. J. (2004) Selective neurotoxic lesions of basolateral and central nuclei of the amygdala produce differential effects on fear conditioning. Journal of Neuroscience 24: 7654–62.CrossRefGoogle ScholarPubMed
Kordower, J. H., Bartus, R. T., Marciano, F. F. and Gash, D. M. (1989) Telencephalic cholinergic system of the new world monkey (Cebus apella): Morphological and cytoarchitectonic assessment and analysis of the projection to the amygdala. Journal of Comparative Neurology 279: 528–45.CrossRefGoogle ScholarPubMed
Kosaka, T., Kosaka, K., Hataguchi, Y.et al. (1987) Catecholaminergic neurons containing GABA-like and/or glutamic acid decarboxylase-like immunoreactivities in various brain regions of the rat. Experimental Brain Research 66: 191–210.CrossRefGoogle ScholarPubMed
Kosofsky, B. E. and Molliver, M. E. (1987) The serotoninergic innervation of cerebral cortex: different classes of axon terminals arise from dorsal and median raphe nuclei. Synapse 1: 153–68.CrossRefGoogle ScholarPubMed
Krettek, J. E. and Price, J. L. (1977a) Projections from the amygdaloid complex to the cerebral cortex and thalamus in the rat and cat. Journal of Comparative Neurology 172: 687–722.CrossRefGoogle Scholar
Krettek, J. E. and Price, J. L. (1977b) Projections from the amygdaloid complex and adjacent olfactory structures to the entorhinal cortex and to the subiculum in the rat and cat. Journal of Comparative Neurology 172: 723–52.CrossRefGoogle Scholar
Krettek, J. E. and Price, J. L. (1978a) Amygdaloid projections to subcortical structures within the basal forebrain and brainstem in the rat and cat. Journal of Comparative Neurology 178: 225–54.CrossRefGoogle Scholar
Krettek, J. E. and Price, J. L. (1978b) A description of the amygdaloid complex in the rat and cat with observations on intra-amygdaloid axonal connections. Journal of Comparative Neurology 178: 255–80.CrossRefGoogle Scholar
Kroner, S., Rosenkranz, J. A., Grace, A. A. and Barrionuevo, G. (2005) Dopamine modulates excitability of basolateral amygdala neurons in vitro. Journal of Neurophysiology 93: 1598–610.CrossRefGoogle ScholarPubMed
Kudo, M. and Niimi, K. (1980) Ascending projections of the inferior colliculus in the cat: an autoradiographic study. Journal of Comparative Neurology 191: 545–56.CrossRefGoogle Scholar
Kudo, M., Itoh, K., Kawamura, S. and Mizuno, N. (1983) Direct projections to the pretectum and the midbrain reticular formation from auditory relay nuclei in the lower brainstem of the cat. Brain Research 288: 13–19.CrossRefGoogle ScholarPubMed
Kudo, M., Tashiro, T., Higo, S., Matsuyama, T. and Kawamura, S. (1984) Ascending projections from the nucleus of the brachium of the inferior colliculus in the cat. Experimental Brain Research 54: 203–11.CrossRefGoogle ScholarPubMed
Kumar, S. S. and Huguenard, J. R. (2001) Properties of excitatory synaptic connections mediated by the corpus callosum in the developing neocortex. Journal of Neurophysiology 86: 2973–85.CrossRefGoogle Scholar
Kuroda, M. and Price, J. L. (1991) Synaptic organization of projections from basal forebrain structures to the mediodorsal thalamic nucleus of the rat. Journal of Comparative Neurology 303: 513–33.CrossRefGoogle ScholarPubMed
LaBar, K. S., LeDoux, J. E., Spencer, D. D. and Phelps, E. A. (1995) Impaired fear conditioning following unilateral temporal lobectomy in humans. Journal of Neuroscience 15: 6846–55.CrossRefGoogle ScholarPubMed
Lamprecht, R., Farb, C. R. and LeDoux, J. E. (2002) Fear memory formation involves p190 RhoGAP and ROCK proteins through a GRB2-mediated complex. Neuron 36: 727–38.CrossRefGoogle ScholarPubMed
Lancel, M., Riezen, H., and Glatt, A. (1992) The time course of sigma activity and slow-wave activity during NREMS in cortical and thalamic EEG of the cat during baseline and after 12 hours of wakefulness. Brain Research 596: 285–95.CrossRefGoogle ScholarPubMed
Landisman, C. E., Long, M. A., Beierlein, M.et al. (2002) Electrical synapses in the thalamic reticular nucleus. Journal of Neuroscience 22: 1002–9.CrossRefGoogle ScholarPubMed
Lang, E. J. and Paré, D. (1997a) Similar inhibitory processes dominate the responses of cat lateral amygdaloid projection neurons to their various afferents. Journal of Neurophysiology 77: 341–52.CrossRefGoogle Scholar
Lang, E. J. and Paré, D. (1997b) Synaptic and synaptically activated intrinsic conductances underlie inhibitory potentials in cat lateral amygdaloid projection neurons in vivo. Journal of Neurophysiology 77: 353–63.CrossRefGoogle Scholar
Lang, E. J. and Paré, D. (1998) Synaptic responsiveness of interneurons of the cat lateral amygdaloid nucleus. Neuroscience 83: 877–89.CrossRefGoogle ScholarPubMed
Larkum, M. E., Zhu, J. J. and Sakman, B. (1999) A new cellular mechanism for coupling inputs arriving at different cortical layers. Nature 398: 338–41.CrossRefGoogle ScholarPubMed
Laroche, S., Jay, T. M. and Thierry, A. M. (1990) Long-term potentiation in the prefrontal cortex following stimulation of the hippocampal CA1/subicular region. Neuroscience Letters 114: 184–90.CrossRefGoogle ScholarPubMed
Laroche, S., Doyère, V., Redini-Del Negro, C. and Burette, F. (1995) Neural mechanisms of associative memory: role of long-term potentiation. In Brain and Memory: Modulation and Mediation of Neuroplasticity, ed. McGaugh, J. L., Weinberger, N. M. and Lynch, G., pp. 277–302. New York: Oxford University Press.CrossRefGoogle Scholar
Larson, J. and Lynch, G. (1986) Role of N-methyl-D aspartate receptors in the induction of synaptic potentiation by burst stimulation patterned after the hippocampal theta rhythm. Brain Research 441: 111–18.CrossRefGoogle Scholar
Larson, J., Wong, D. and Lynch, G. (1986) Patterned stimulation at the theta frequency is optimal for the induction of hippocampal long-term potentiation. Brain Research 368: 347–50.CrossRefGoogle ScholarPubMed
Lavin, A. and Grace, A. A. (1994) Modulation of dorsal thalamic cell activity by the ventral pallidum: its role in the regulation of thalamocortical activity by the basal ganglia. Synapse 18: 104–27.CrossRefGoogle ScholarPubMed
Lavoie, B. and Parent, A. (1994) Pedunculopontine nucleus in the squirrel monkey: distribution of cholinergic and monoaminergic neurons in the mesopontine tegmentum with evidence for the presence of glutamate in cholinergic neurons. Journal of Comparative Neurology 344: 190–209.CrossRefGoogle ScholarPubMed
Law-Tho, D., Desce, J. M. and Crépel, F. (1995) Dopamine favours the emergence of long-term depression versus long-term potentiation in slices of rat prefrontal cortex. Neuroscience Letters 188: 125–8.CrossRefGoogle ScholarPubMed
LeDoux, J. E. (2000a) The amygdala and emotion: a view through fear. In The Amygdala: A Functional Analysis, ed. Aggleton, J. P., pp. 289–310. Oxford: Oxford University Press.Google Scholar
LeDoux, J. E. (2000b) Emotion circuits in the brain. Annual Review of Neuroscience 23: 155–84.CrossRefGoogle Scholar
LeDoux, J. E., Sakaguchi, A. and Reis, D. J. (1984) Subcortical efferent projections of the medial geniculate nucleus mediate emotional responses conditioned to acoustic stimuli. Journal of Neuroscience 4: 683–98.CrossRefGoogle ScholarPubMed
LeDoux, J. E., Ruggiero, D. A., Forest, R., Stornetta, R. and Reis, D. J. (1987) Topographic organization of convergent projections to the thalamus from the inferior colliculus and spinal cord in the rat. Journal of Comparative Neurology 264: 123–46.CrossRefGoogle ScholarPubMed
LeDoux, J. E., Iwata, J., Cicchetti, P. and Reis, D. J. (1988) Different projections of the central amygdaloid nucleus mediate autonomic and behavioral correlates of conditioned fear. Journal of Neuroscience 8: 2517–29.CrossRefGoogle ScholarPubMed
LeDoux, J. E., Cicchetti, P., Xagoraris, A. and Romanski, L. M. (1990a) The lateral amygdaloid nucleus: sensory interface of the amygdala in fear conditioning. Journal of Neuroscience 10: 1062–9.CrossRefGoogle Scholar
LeDoux, J. E., Farb, C. and Ruggiero, D. A. (1990b) Topographic organization of neurons in the acoustic thalamus that project to the amygdala. Journal of Neuroscience 10: 1043–54.CrossRefGoogle Scholar
Lee, H. and Kim, J. J. (1998) Amygdalar NMDA receptors are critical for new fear learning in previously fear-conditioned rats. Journal of Neuroscience 18: 8444–54.CrossRefGoogle ScholarPubMed
Lee, H. K., Barbarosie, M., Kameyama, K., Bear, M. F. and Huganir, R. L. (2000) Regulation of distinct AMPA receptor phosphorylation sites during bidirectional synaptic plasticity. Nature 405: 955–9.CrossRefGoogle ScholarPubMed
Lee, K. H. and McCormick, D. A. (1995) Acetylcholine excites GABAergic neurons of the ferret perigeniculate nucleus through nicotinic receptors. Journal of Neurophysiology 73: 2123–8.CrossRefGoogle ScholarPubMed
Lee, K. and McCormick, D. A. (1996) Abolition of spindle oscillations by serotonin and norepinephrine in the ferret lateral geniculate and perigeniculate nuclei in vitro. Neuron 17: 309–21.CrossRefGoogle ScholarPubMed
Lee, M. G., Hassani, O. K., Alonso, A. and Jones, B. E. (2005) Cholinergic basal forebrain neurons burst with theta during waking and paradoxical sleep. Journal of Neuroscience 25: 4365–9.CrossRefGoogle ScholarPubMed
Lee, M. S., Friedberg, M. H. and Ebner, F. F. (1994) The role of GABA-mediated inhibition in the rat ventral posterior medial thalamus. I. Assessment of receptive field changes following thalamic reticular nucleus lesions. Journal of Neurophysiology 71: 1702–15.CrossRefGoogle ScholarPubMed
Lenz, F. A. and Dougherty, P. M. (1997) Pain processing in the human thalamus. In Thalamus, vol. 2 (Experimental and Clinical Aspects), ed. Steriade, M., Jones, E. G. and McCormick, D. A., pp. 617–51. Amsterdam: Elsevier.
Leonard, B. W., Amaral, D. G., Squire, L. R. and Zola-Morgan, S. (1995) Transient memory impairment in monkeys with bilateral lesions of the entorhinal cortex. Journal of Neuroscience 15: 5367–659.CrossRefGoogle ScholarPubMed
Leonard, C. S. and Llinás, R. R. (1990) Electrophysiology of mammalian pedunculopontine and laterodorsal tegmental neurons in vitro: implications for the control of REM sleep. In Brain Cholinergic Systems, ed. Steriade, M. and Biesold, D., pp. 205–23. Oxford: Oxford University Press.
Leonard, C. S. and Llinás, R. R. (1994) Serotonergic and cholinergic inhibition of mesopontine cholinergic neurons controlling REM sleep: an in vitro electrophysiological study. Neuroscience 59: 309–30.CrossRefGoogle Scholar
Leonard, C. S., Kerman, I., Blaha, G., Taveras, E., and Taylor, B. (1995a) Interdigitation of nitric oxyde synthase-, tyrosine hydroxylase- and serotonin-containing neurons in and around the laterodorsal and pedunculopontine tegmental nuclei of the guinea pig. Journal of Comparative Neurology 362: 411–32.CrossRefGoogle Scholar
Leonard, C. S., Rao, S. and Sanchez, R. M. (1995b) Patterns of neuromodulation and the nitric oxide signaling pathway in mesopontine cholinergic neurons. Seminars in the Neurosciences 7: 319–28.CrossRefGoogle Scholar
Leonard, C. S., Rao, S. R. and Inoue, T. (2000) Serotonergic inhibition of action potential evoked calcium transients in NOS-containing mesopontine cholinergic neurons. Journal of Neurophysiology 84: 1558–72.CrossRefGoogle ScholarPubMed
Leopold, D. A., Murayama, Y. and Logothetis, N. K. (2003) Very slow activity fluctuations in monkey visual cortex: implications for functional brain imaging. Cerebral Cortex 13: 422–33.CrossRefGoogle ScholarPubMed
Leresche, N., Jassik-Gerschenfeld, D., Haby, M., Soltesz, I. and Crunelli, V. (1990) Pacemaker-like and other types of spontaneous membrane potential oscillations of thalamocortical cells. Neuroscience Letters 113: 72–7.CrossRefGoogle ScholarPubMed
Leresche, N., Lightowler, S., Soltesz, I., Jassik-Gerschenfeld, D. and Crunelli, V. (1991) Low-frequency oscillatory activities intrinsic to rat and cat thalamocortical cells. Journal of Physiology (London) 441: 155–74.CrossRefGoogle ScholarPubMed
Leung, L. S. (1984) Model of gradual phase shift of theta rhythm in the rat. Journal of Neurophysiology 52: 1051–65.CrossRefGoogle ScholarPubMed
LeVay, S. (1973) Synaptic patterns in the visual cortex of the cat and monkey. Electron microscopy of Golgi preparations. Journal of Comparative Neurology 150: 53–86.CrossRefGoogle ScholarPubMed
Lévesque, M., Charara, A., Gagnon, S., Parent, A. and Deschênes, M. (1996) Corticostriatal projections from layer V cells in rat are collaterals of long-range corticofugal axons. Brain Research 709: 311–15.CrossRefGoogle Scholar
Levitt, P. and Moore, R. Y. (1978) Noradrenaline neuron innervation of the neocortex in the rat. Brain Research 139: 219–31.CrossRefGoogle ScholarPubMed
Lhermitte, F., Gautier, J. C., Marteau, R. and Chain, F. (1963) Consciousness disorders and akinetic mutism. Anatomo-clinical study of a bilateral paramedian softening of the cerebral peduncle and thalamus. Revue Neurologique (Paris) 109: 115–31.Google ScholarPubMed
Li, L., Miller, E. K. and Desimone, R. (1993) The representation of stimulus familiarity in anterior inferior temporal cortex. Journal of Neurophysiology 69: 1918–29.CrossRefGoogle ScholarPubMed
Li, R., Nishijo, H., Wang, Q. X.et al. (2001) Light and electron microscopic study of cholinergic and noradrenergic elements in the basolateral nucleus of the rat amygdala: Evidence for interactions between the two systems. Journal of Comparative Neurology 439: 411–25.CrossRefGoogle ScholarPubMed
Li, R., Nishijo, H., Ono, T., Ohtani, Y. and Ohtani, O. (2002) Synapses on GABAergic neurons in the basolateral nucleus of the rat amygdala: double-labeling immunoelectron microscopy. Synapse 43: 42–50.CrossRefGoogle ScholarPubMed
Li, Y. Q., Jia, H. G., Rao, Z. R. and Shi, J. W. (1990) Serotonin-, substance P- or leucine-enkephalin-containing neurons in the midbrain periaqueductal gray and nucleus raphe dorsalis send projection fibers to the central amygdaloid nucleus in the rat. Neuroscience Letters 120: 124–7.CrossRefGoogle ScholarPubMed
Liang, K. C. and McGaugh, J. L. (1983) Lesions of the stria terminals attenuate the enhancing effect of post-training epinephrine on retention of an inhibitory avoidance response. Behavioral Brain Research 9: 49–58.CrossRefGoogle Scholar
Liang, K. C., McGaugh, J. L. and Yao, H. Y. (1990) Involvement of amygdala pathways in the influence of post-training intra-amygdala norepinephrine and peripheral epinephrine on memory storage. Brain Research 508: 225–33.CrossRefGoogle ScholarPubMed
Likhtik, E., Pelletier, J. G., Paz, R. and Paré, D. (2005) Prefrontal control of the amygdala. Journal of Neuroscience 25: 7429–37.CrossRefGoogle ScholarPubMed
Lin, J. S., Sakai, K. and Jouvet, M. (1988) Evidence for histaminergic arousal mechanisms in the hypothalamus of cats. Neuropharmacology 27: 111–22.CrossRefGoogle Scholar
Lin, J. S., Sakai, K., Vanni-Mercier, G. and Jouvet, M. (1989) A critical role of the posterior hypothalamus in the mechanisms of wakefulness determined by microinjections of muscimol in freely moving cats. Brain Research 479: 225–40.CrossRefGoogle ScholarPubMed
Lin, J. S., Hou, Y., Sakai, K. and Jouvet, M. (1996) Histaminergic descending inputs to the mesopontine tegmentum and their role in the control of cortical activation and wakefulness in the cat. Journal of Neuroscience 16: 1523–37.CrossRefGoogle ScholarPubMed
Lindström, S. and Wróbel, A. (1990) Frequency dependent corticofugal excitation of principal cells in the cat's dorsal lateral geniculate nucleus. Experimental Brain Research 79: 313–8.CrossRefGoogle ScholarPubMed
Lindvall, O., Bjorklund, A. and Divac, I. (1978) Organization of catecholamine neurons projecting to the frontal cortex in the rat. Brain Research 142: 1–24.CrossRefGoogle ScholarPubMed
Linke, R., Braune, G. and Schwegler, H. (2000) Differential projection of the posterior paralaminar thalamic nuclei to the amygdaloid complex in the rat. Experimental Brain Research 134: 520–32.CrossRefGoogle ScholarPubMed
Liu, X. B. and Jones, E. G. (1999) Predominance of corticothalamic synaptic inputs to thalamic reticular nucleus neurons in the rat. Journal of Comparative Neurology 414: 67–79.3.0.CO;2-Z>CrossRefGoogle ScholarPubMed
Liu, X. B., Warren, R. A. and Jones, E. G. (1995) Synaptic distribution of afferents from reticular nucleus in ventroposterior nucleus of cat thalamus. Journal of Comparative Neurology 352: 187–202.CrossRefGoogle ScholarPubMed
Liubashina, O., Jolkkonen, E. and Pitkänen, A. (2000) Projections from the central nucleus of the amygdala to the gastric related area of the dorsal vagal complex: a Phaseolus vulgaris-leucoagglutinin study in rat. Neuroscience Letters 291: 85–8.CrossRefGoogle ScholarPubMed
Livingstone, M. S. and Hubel, D. H. (1981) Effects of sleep and arousal on the processing of visual information in the cat. Nature 291: 554–61.CrossRefGoogle ScholarPubMed
Llinás, R. R. (1988) The intrinsic electrophysiological properties of mammalian neurons: insights into central nervous system function. Science 242: 1654–64.CrossRefGoogle ScholarPubMed
Llinás, R. R. and Jahnsen, H. (1982) Electrophysiology of mammalian thalamic neurones in vitro. Nature 297: 406–8.CrossRefGoogle ScholarPubMed
Llinás, R., and Ribary, U. 1992. Rostrocaudal scan in human brain: a global characteristic for the 40 Hz response during sensory input. In Induced Rhythms in the Brain, ed. Basar, E. and Bullock, T., pp. 147–54. Boston, MA: Birkhaüser.
Llinás, R. R. and Ribary, U. (1993) Coherent 40-Hz oscillation characterizes dream state in humans. Proceedings of the National Academy of Sciences USA 90: 2078–81.CrossRefGoogle ScholarPubMed
Llinás, R., Grace, A. A. and Yarom, Y. (1991) In vitro neurons in mammalian cortical layer 4 exhibit intrinsic oscillatory activity in the 10- to 50-Hz frequency range. Proceedings of the National Academy of Sciences USA 88: 897–901.CrossRefGoogle ScholarPubMed
Llinás, R. R., Ribary, U., Jeanmonod, D., Kronberg, E. and Mitra, P. P. (1999) Thalamocortical dysrhythmia: a neurological and neuropsychiatric syndrome characterized by magnetoencephalography. Proceedings of the National Academy of Sciences USA 96: 15222–7.CrossRefGoogle ScholarPubMed
Llinás, R. R., Leznik, E. and Urbano, F. J. (2002) Temporal binding via cortical coincidence detection of specific and nonspecific thalamocortical inputs: a voltage-dependent dye-imaging study in mouse brain slices. Proceedings of the National Academy of Sciences USA 99: 449–54.CrossRefGoogle ScholarPubMed
Llinás, R. R., Urbano, F. J., Leznik, E., Ramirez, R. R. and Marle, H. H. F. (2005) Rhythmic and dysrhythmic thalamocortical dynamics: GABA systems and the edge effect. Trends in Neurosciences 28(6): 325–33.CrossRefGoogle ScholarPubMed
Logothetis, N. K. (2003) The underpinnings of the BOLD functional magnetic resonance imaging signal. Journal of Neuroscience 23: 3963–71.CrossRefGoogle ScholarPubMed
Long, M. A., Landisman, C. E. and Connors, B. W. (2004) Small clusters of electrically coupled neurons generate synchronous rhythms in the thalamic reticular nucleus. Journal of Neuroscience 24: 341–9.CrossRefGoogle ScholarPubMed
Loomis, A. L., Harvey, N. and Hobart, G. A. (1938) Distribution of disturbance patterns in the human electroencephalogram, with special reference to sleep. Journal of Neurophysiology 1: 413–30.CrossRefGoogle Scholar
Lopes da Silva, F. H., Rotterdam, A., Storm van Leeuwen, W. and Tielen, A. M. (1970) Dynamic characteristics of visual evoked potentials in the dog. II. Beta frequency selectivity in evoked potentials and background activity. Electroencephalography and Clinical Neurophysiology 29: 260–8.CrossRefGoogle ScholarPubMed
Lopes da Silva, F. H. and Storm van Leeuwen, W. (1978) The cortical alpha rhythm in dog: depth and surface profile of phase. In Architecture of the Cerebral Cortex (IBRO monograph series, vol. 3), ed. Brazier, M. A. B. and Petsche, H., pp. 319–33. New York: Raven Press.
Lopes da Silva, F. H., Lierop, T. H. M. T., Schrijer, C. F. M. and Storm Van Leeuwen, W. (1973) Organization of thalamic and cortical alpha rhythm: spectra and coherences. Electroencephalography and Clinical Neurophysiology 35: 627–39.CrossRefGoogle Scholar
Lopes da Silva, F. H., Vos, J. E., Mooibroeck, J. and Rotterdam, A. (1980) Relative contribution of intracortical and thalamo-cortical processes in the generation of alpha rhythms, revealed by partial coherence analysis. Electroencephalography and Clinical Neurophysiology 50: 449–56.CrossRefGoogle Scholar
Lopez De Armentia, M. and Sah, P. (2004) Firing properties and connectivity of neurons in the rat lateral central nucleus of the amygdala. Journal of Neurophysiology 92: 1285–94.CrossRefGoogle ScholarPubMed
Lorente de No, R. (1933) Studies on the structure of the cerebral cortex. I. The area entorhinalis. Journal für Psychologie und Neurologie 45: 381–438.Google Scholar
Loretan, K., Bissière, S. and Lüthi, A. (2004) Dopaminergic modulation of spontaneous inhibitory network activity in the lateral amygdala. Neuropharmacology 47: 631–9.CrossRefGoogle ScholarPubMed
Losier, B. J. and Semba, K. (1993) Dual projections of single cholinergic and aminergic brainstem neurons to the thalamus and basal forebrain in the rat. Brain Research 604: 41–52.CrossRefGoogle ScholarPubMed
Loughlin, S. E. and Fallon, J. H. (1983) Dopaminergic and non-dopaminergic projections to amygdala from substantia nigra and ventral tegmental area. Brain Research 262: 334–8.CrossRefGoogle ScholarPubMed
Lu, J., Greco, M. A., Shiromani, P. and Saper, C. B. (2000) Effect of lesions of the ventrolateral preoptic nucleus on NREM and REM sleep. Journal of Neuroscience 20: 3830–42.CrossRefGoogle ScholarPubMed
Lu, K. T., Walker, D. L. and Davis, M. (2001) Mitogen-activated protein kinase cascade in the basolateral nucleus of amygdala is involved in extinction of fear-potentiated startle. Journal of Neuroscience 21: NIL17–21.CrossRefGoogle ScholarPubMed
Lu, W. Y., Man, H. Y., Ju, W.et al. (2001) Activation of synaptic NMDA receptors induces membrane insertion of new AMPA receptors and LTP in cultured hippocampal neurons. Neuron 29: 243–54.CrossRefGoogle ScholarPubMed
Lucretius, T. C. (1988) On the Nature of the Universe (translated by R. E. Latham). London: Penguin Books.Google Scholar
Luebke, J. I., Greene, R. W., Semba, K.et al. (1992) Serotonin hyperpolarizes cholinergic low-threshold burst neurons in the rat laterodorsal tegmental nucleus in vitro. Proceedings of the National Academy of Sciences USA 89: 743–7.CrossRefGoogle ScholarPubMed
Lugaresi, E., Medori, R.Montagna, P.et al. (1986) Fatal familial insomnia and dysautonomia with selective degeneration of thalamic nuclei. New England Journal of Medicine 315: 997–1003.CrossRefGoogle ScholarPubMed
Luiten, P. G., Spencer, D. G., Traber, J. and Gaykema, R. P. (1985) The pattern of cortical projections from the intermediate parts of the magnocellular nucleus basalis in the rat demonstrated by tracing with Phaseolus vulgaris-leucoagglutinin. Neuroscience Letters 57: 137–42.CrossRefGoogle ScholarPubMed
Luiten, P. G., Gaykema, R. P., Traber, J. and Spencer, D. G. (1987) Cortical projection patterns of magnocellular basal nucleus subdivisions as revealed by anterogradely transported Phaseolus vulgaris leucoagglutinin. Brain Research 413: 229–50.CrossRefGoogle ScholarPubMed
Lund, J. S. (1973) Organization of neurons in the visual cortex, area 17, of the monkey (Macaca mulatta). Journal of Comparative Neurology 147: 455–96.CrossRefGoogle Scholar
Lüthi, A. and McCormick, D. A. (1998) Periodicity of thalamic synchronized oscillations: the role of Ca2+-mediated upregulation of Ih. Neuron 20: 553–63.CrossRefGoogle ScholarPubMed
Lux, H. D. and Neher, E. (1973) The equilibrium time course of [K+]o in cat cortex. Experimental Brain Research 17: 190–205.CrossRefGoogle Scholar
Lytton, W. W. and Sejnowski, T. J. (1991) Simulation of cortical pyramidal neurons synchronized by inhibitory interneurons. Journal of Neurophysiology 66: 1059–79.CrossRefGoogle ScholarPubMed
Lytton, W. W., Contreras, D., Destexhe, A. and Steriade, M. (1997) Dynamic interactions determine partial thalamic quiescence in a computer network model of spike-and-wave seizures. Journal of Neurophysiology 77: 1679–96.CrossRefGoogle Scholar
Ma, Q. P., Yin, G. F., Ai, M. K. and Han, J. S. (1991) Serotonergic projections from the nucleus raphe dorsalis to the amygdala in the rat. Neuroscience Letters 134: 21–4.CrossRefGoogle ScholarPubMed
Macchi, G., Rossi, G., Abbamondi, A. L.et al. (1997) Diffuse thalamic degeneration in fatal familial insomnia. A morphometric study. Brain Research 771: 154–8.CrossRefGoogle ScholarPubMed
Mackay, A. V. P. (1998) Letter to the Editor. Trends in Neurosciences 21: 146.CrossRefGoogle Scholar
Magill, P. J., Bolam, P. and Bevan, M. D. (2000) Relationship of activity in the subthalamic nucleus – globus pallidus network to cortical EEG. Journal of Neuroscience 20: 820–33.CrossRefGoogle Scholar
Mahanty, N. K. and Sah, P. (1998) Calcium-permeable AMPA receptors mediate long-term potentiation in interneurons in the amygdala. Nature 394: 683–7.CrossRefGoogle ScholarPubMed
Mahon, S., Deniau, J. M. and Charpier, S. (2001) Relationship between EEG potentials and intracellular activity of striatal and cortico-striatal neurons: an in vivo study under different anesthetics. Cerebral Cortex 11: 360–73.CrossRefGoogle Scholar
Mainen, Z. F. and Sejnowski, T. J. (1995) Reliability of spike timing in neocortical neurons. Science 268: 1503–6.CrossRefGoogle ScholarPubMed
Malenka, R. C. and Nicoll, R. A. (1993) NMDA-receptor-dependent synaptic plasticity: multiple forms and mechanisms. Trends in Neurosciences 16: 521–7.CrossRefGoogle ScholarPubMed
Maloney, K. J., Cape, E. G., Gotman, J. and Jones, B. E. (1997) High-frequency gamma electroencephalogram activity in association with sleep-wake states and spontaneous behaviors in the rat. Neuroscience 76: 541–55.CrossRefGoogle ScholarPubMed
Maltais, S., Coté, S., Drolet, G. and Falardeau, P. (2000) Cellular colocalization of dopamine D1 mRNA and D2 receptor in rat brain using a dopamine receptor specific polyclonal antibody. Progress in Neuro-Psychopharmacology and Biological Psychiatry 24: 1127–49.CrossRefGoogle ScholarPubMed
Manaye, K. F., Zweig, R., Wu, D.et al. (1999) Quantification of cholinergic and select non-cholinergic mesopontine neuronal populations in the human brain. Neuroscience 89: 759–70.CrossRefGoogle ScholarPubMed
Manetto, V., Medori, R., Cortelli, P.et al. (1992) Fatal familial insomnia: clinical and pathological study of five new cases. Neurology 42: 312–19.CrossRefGoogle ScholarPubMed
Manns, I. D., Mainville, L. and Jones, B. E. (2001) Evidence for glutamate, in addition to acetylcholine and GABA, neurotransmitter synthesis in basal forebrain neurons projecting to the entorhinal cortex. Neuroscience 107: 249–63.CrossRefGoogle ScholarPubMed
Mao, B. Q., Hamzei-Sichani, F., Aronov, D., Froemke, R. C. and Yuste, R. (2003) Dynamics of spontaneous activity in neocortical slices. Neuron 32: 883–98.CrossRefGoogle Scholar
Maquet, P., Péters, J. M., Aerts, J.et al. (1996) Functional neuroanatomy of human rapid-eye-movement sleep and dreaming. Nature 383: 163–6.CrossRefGoogle ScholarPubMed
Maquet, P., Degueldre, C., Delfiore, G.et al. (1997) Functional neuroanatomy of human slow wave sleep. Journal of Neuroscience 17: 2807–12.CrossRefGoogle ScholarPubMed
Maquet, P., Laureys, S., Peigneux, P.et al. (2000) Experience-dependent changes in cerebral activation during human REM sleep. Nature Neuroscience 3: 831–6.CrossRefGoogle ScholarPubMed
Maren, S. and Fanselow, M. S. (1995) Synaptic plasticity in the basolateral amygdala induced by hippocampal formation stimulation in vivo. Journal of Neuroscience 15: 7548–64.CrossRefGoogle ScholarPubMed
Maren, S. and Quirk, G. J. (2004) Neuronal signaling of fear memory. Nature Reviews Neuroscience 5: 844–52.CrossRefGoogle ScholarPubMed
Markram, H., Lübke, J., Frotscher, M., Roth, A. and Sakmann, B. (1997) Physiology and anatomy of tufted pyramidal neurons in the developing rat neocortex. Journal of Physiology (London) 500: 409–40.CrossRefGoogle ScholarPubMed
Markram, H., Wang, M. and Tsodycs, M. (1998) Differential signaling via the same axon of neocortical pyramidal neurons. Proceedings of the National Academy of Sciences USA 95: 5323–8.CrossRefGoogle ScholarPubMed
Markram, H., Toledo-Rodriguez, M., Wang, Y.et al. (2004) Interneurons of the neocortical inhibitory system. Nature Reviews Neuroscience 5: 793–807.CrossRefGoogle ScholarPubMed
Marshall, L., Mölle, M. and Born, J. (2003) Spindle and slow wave rhythms at slow wave sleep transitions are linked to strong shifts in the cortical direct current potential. Neuroscience 121: 1047–53.CrossRefGoogle ScholarPubMed
Martin, J. J. (1997) Degenerative diseases of the human thalamus. In Thalamus, vol. 2 (Experimental and Clinical Aspects), ed. Steriade, M., Jones, E. G. and McCormick, D. A., pp. 653–79. Oxford: Elsevier.Google Scholar
Martin, J. J., Yap, M., Nei, I. P. and Tan, T. E. (1983) Selective thalamic degeneration – report of a case with memory and mental disturbances. Clinical Neuropathology 2: 156–62.Google ScholarPubMed
Martina, M., Royer, S. and Paré, D. (1999) Physiological properties of central medial and central lateral amygdala neurons. Journal of Neurophysiology 82: 1843–54.CrossRefGoogle ScholarPubMed
Martina, M., Royer, S. and Paré, D. (2001) Propagation of neocortical inputs in the perirhinal cortex. Journal of Neuroscience 21: 2878–88.CrossRefGoogle ScholarPubMed
Mascagni, F., McDonald, A. J. and Coleman, J. R. (1993) Corticoamygdaloid and corticocortical projections of the rat temporal cortex: A Phaseolus vulgaris leucoagglutinin study. Neuroscience 57: 697–715.CrossRefGoogle ScholarPubMed
Massimini, M. and Amzic, F. (2001) Extracellular calcium fluctuations and intracellular potentials in the cortex during the slow sleep oscillation. Journal of Neurophysiology 85: 1346–50.CrossRefGoogle ScholarPubMed
Massimini, M., Rosanova, M. and Mariotti, M. (2003) EEG slow (∼1 Hz) waves are associated with nonstationarity of thalamo-cortical sensory processing in the sleeping human. Journal of Neurophysiology 89: 1205–13.CrossRefGoogle ScholarPubMed
Massimini, M., Huber, R., Ferrarelli, F. and Tononi, G. (2004) The sleep slow oscillation as a traveling wave. Journal of Neuroscience 24: 6862–70.CrossRefGoogle ScholarPubMed
McCarley, R. W. and Hobson, J. A. (1975) Neuronal excitability modulation over the sleep cycle: a structural and mathematical model. Science 189: 58–60.CrossRefGoogle ScholarPubMed
McCarley, R. W., Nelson, J. P. and Hobson, J. A. (1978) Ponto-geniculo-occipital (PGO) burst neurons: correlative evidence for neuronal generators of PGO waves. Science 201: 269–72.CrossRefGoogle ScholarPubMed
McCarley, R. W., Benoit, O. and Barrionuevo, G. (1983) Lateral geniculate nucleus unitary discharge in sleep and waking: state- and rate-specific aspects. Journal of Neurophysiology 50: 798–818.CrossRefGoogle ScholarPubMed
McCormick, D. A. (1992) Neurotransmitter actions in the thalamus and cerebral cortex and their role in neuromodulation of thalamocortical activity. Progress in Neurobiology 39: 337–88.CrossRefGoogle ScholarPubMed
McCormick, D. A. and Pape, H. C. (1988) Acetylcholine inhibits identified interneurons in the cat lateral geniculate nucleus. Nature 334: 246–8.CrossRefGoogle ScholarPubMed
McCormick, D. A. and Pape, H. C. (1990a) Properties of a hyperpolarization-activated cation current and its role in rhythmic oscillation in thalamic relay neurones. Journal of Physiology (London) 431: 291–318.CrossRefGoogle Scholar
McCormick, D. A. and Pape, H. C. (1990b) Noradrenergic and serotonergic modulation of a hyperpolarization-activated cation current in thalamic relay cells. Journal of Physiology (London) 431: 319–42.CrossRefGoogle Scholar
McCormick, D. A. and Wang, Z. (1991) Serotonin and noradrenaline excite GABAergic neurones of the guinea-pig and cat nucleus reticularis thalami. Journal of Physiology (London) 442: 235–55.CrossRefGoogle ScholarPubMed
McCormick, D. A. and Williamson, A. (1989) Convergence and divergence of neurotransmitter action in human cerebral cortex. Proceedings of the National Academy of Sciences USA 86: 8098–102.CrossRefGoogle ScholarPubMed
McCormick, D. A. and Williamson, A. (1991) Modulation of neuronal firing mode in cat and guinea-pig LGNd by histamine: possible cellular mechanisms of histaminergic control of arousal. Journal of Neuroscience 11: 3188–99.CrossRefGoogle ScholarPubMed
McCormick, D. A., Connors, B. W., Lighthall, J. W. and Prince, D. A. (1985) Comparative electrophysiology of pyramidal and sparsely spiny stellate neurons of the neocortex. Journal of Neurophysiology 54: 782–806.CrossRefGoogle ScholarPubMed
McDonald, A. J. (1982) Cytoarchitecture of the central amygdaloid nucleus of the rat. Journal of Comparative Neurology 208: 401–18.CrossRefGoogle ScholarPubMed
McDonald, A. J. (1985) Immunohistochemical identification of gamma-aminobutyric acid-containing neurons in the rat basolateral amygdala. Neuroscience Letters 53: 203–7.CrossRefGoogle ScholarPubMed
McDonald, A. J. (1992a) Cell types and intrinsic connections of the amygdala. In The Amygdala: Neurobiological Aspects of Emotion, Memory, and Mental Dysfunction, ed. Aggleton, J. P., pp. 67–96. New York: Wiley-Liss.Google Scholar
McDonald, A. J. (1992b) Projection neurons of the basolateral amygdala: A correlative Golgi and retrograde tract tracing study. Brain Research Bulletin 28: 179–85.CrossRefGoogle Scholar
McDonald, A. J. (1998) Cortical pathways to the mammalian amygdala. Progress in Neurobiology 55: 257–332.CrossRefGoogle ScholarPubMed
McDonald, A. J. (2003) Is there an amygdala and how far does it extend? An anatomical perspective. Annals of the New York Academy of Sciences 985: 1–21.CrossRefGoogle ScholarPubMed
McDonald, A. J. and Augustine, J. R. (1993) Localization of GABA-like immunoreactivity in the monkey amygdala. Neuroscience 52: 281–94.CrossRefGoogle ScholarPubMed
McDonald, A. J. and Betette, R. L. (2001) Parvalbumin-containing neurons in the rat basolateral amygdala: morphology and co-localization of Calbindin-D(28k). Neuroscience 102: 413–25.CrossRefGoogle Scholar
McDonald, A. J. and Mascagni, F. (1997) Projections of the lateral entorhinal cortex to the amygdala. Neuroscience 77: 445–59.CrossRefGoogle ScholarPubMed
McDonald, A. J. and Mascagni, F. (2001a) Colocalization of calcium-binding proteins and GABA in neurons of the rat basolateral amygdala. Neuroscience 105: 681–93.CrossRefGoogle Scholar
McDonald, A. J. and Mascagni, F. (2001b) Localization of the CB1 type cannabinoid receptor in the rat basolateral amygdala: high concentrations in a subpopulation of cholecystokinin-containing interneurons. Neuroscience 107: 641–52.CrossRefGoogle Scholar
McDonald, A. J. and Mascagni, F. (2002).Immunohistochemical characterization of somatostatin containing interneurons in the rat basolateral amygdala. Brain Research 943: 237–44.CrossRefGoogle ScholarPubMed
McDonald, A. J. and Mascagni, F. (2004) Parvalbumin-containing interneurons in the basolateral amygdala express high levels of the alpha1 subunit of the GABA A receptor. Journal of Comparative Neurology 473: 137–46.CrossRefGoogle Scholar
McDonald, A. J., Mascagni, F. and Guo, L. (1996) Projections of the medial and lateral prefrontal cortices to the amygdala: A Phaseolus vulgaris leucoagglutinin study in the rat. Neuroscience 71: 55–75.CrossRefGoogle ScholarPubMed
McDonald, A. J., Muller, J. F. and Mascagni, F. (2002) GABAergic innervation of alpha type II calcium/calmodulin-dependent protein kinase immunoreactive pyramidal neurons in the rat basolateral amygdala. Journal of Comparative Neurology 446: 199–218.CrossRefGoogle ScholarPubMed
McGaugh, J. L. (1966) Time-dependent processes in memory storage. Science 153: 1351–8.CrossRefGoogle ScholarPubMed
McGaugh, J. L. (2002) Memory consolidation and the amygdala: a systems perspective. Trends in Neurosciences 25: 456–61.CrossRefGoogle ScholarPubMed
McGaugh, J. L. and Gold, P. E. (1976) Modulation of memory by electrical stimulation of the brain. In Neural Mechanisms of Learning and Memory, ed. Rosenzweig, M. R. and Bennett, E. L., pp. 549–60. Cambridge, MA: MIT Press.Google Scholar
McGinty, D. J. and Harper, R. M. (1976) Dorsal raphe neurons: depression of firing during sleep in cats. Brain Research 101: 569–75.CrossRefGoogle ScholarPubMed
McKernan, M. G. and Shinnick-Gallagher, P. (1997) Fear conditioning induces a lasting potentiation of synaptic currents in vitro. Nature 390: 607–11.CrossRefGoogle ScholarPubMed
McKinney, M., Coyle, J. T. and Hedreen, J. C. (1983) Topographic analysis of the innervation of the rat neocortex and hippocampus by the basal forebrain cholinergic system. Journal of Comparative Neurology 217: 103–21.CrossRefGoogle ScholarPubMed
Mechawar, N., Cozzari, C. and Descarries, L. (2000) Cholinergic innervation in adult rat cerebral cortex: a quantitative immunocytochemical description. Journal of Comparative Neurology 428: 305–18.3.0.CO;2-Y>CrossRefGoogle ScholarPubMed
Merica, H. and Fortune, R. D. (2003) A unique pattern of sleep structure is found to be identical at all cortical sites: a neurobiological interpretation. Cerebral Cortex 13: 1044–50.CrossRefGoogle ScholarPubMed
Mesulam, M. M., Mufson, E. J., Wainer, B. H. and Levey, A. I. (1983a) Central cholinergic pathways in the rat: an overview based on an alternative nomenclature (Ch1–Ch6). Neuroscience 10: 1185–201.CrossRefGoogle Scholar
Mesulam, M.-M., Mufson, E. J., Levey, A. I. and Wainer, B. H. (1983b) Cholinergic innervation of cortex by the basal forebrain: cytochemistry and cortical connections of the septal area, diagonal band nuclei, nucleus basalis (substantia innominata), and hypothalamus in the rhesus monkey. Journal of Comparative Neurology 214: 170–97.CrossRefGoogle Scholar
Mesulam, M. M., Mufson, E. J., Levey, A. I. and Wainer, B. H. (1984) Atlas of cholinergic neurons in the forebrain and upper brainstem of the macaque based on monoclonal choline acetyltransferase immunohistochemistry and acetylcholinesterase histochemistry. Neuroscience 12: 669–86.CrossRefGoogle ScholarPubMed
Metherate, R., Cox, C. L. and Ashe, J. H. (1992) Cellular bases of neocortical activation: modulation of neural oscillations by the nucleus basalis and endogenous acetylcholine. Journal of Neuroscience 12: 4701–11.CrossRefGoogle ScholarPubMed
Meunier, M., Bachevalier, J., Mishkin, M. and Murray, E. A. (1993) Effects on visual recognition of combined and separate ablations of the entorhinal and perirhinal cortex in rhesus monkeys. Journal of Neuroscience 13: 5418–32.CrossRefGoogle ScholarPubMed
Meunier, M., Hadfield, W., Bachevalier, J. and Murray, E. A. (1996) Effects of rhinal cortex lesions combined with hippocampectomy on visual recognition memory in rhesus monkeys. Journal of Neurophysiology 75: 1190–205.CrossRefGoogle ScholarPubMed
Miettinen, N., Koivisto, E., Riekkinen, P. and Miettinen, R. (1996) Co-existence of parvalbumin and GABA in non-pyramidal neurons of the rat entorhinal cortex. Brain Research 706: 113–22.CrossRefGoogle Scholar
Miettinen, M., Pitkänen, A. and Miettinen, R. (1997) Distribution of calretinin-immunoreactivity in the rat entorhinal cortex: coexistence with GABA. Journal of Comparative Neurology 378: 363–78.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Milad, M. R. and Quirk, G. J. (2002) Neurons in medial prefrontal cortex signal memory for fear extinction. Nature 420: 70–4.CrossRefGoogle ScholarPubMed
Miller, E. K., Li, L. and Desimone, R. (1993) Activity of neurons in anterior inferior temporal cortex during a short-term memory task. Journal of Neuroscience 13: 1460–78.CrossRefGoogle ScholarPubMed
Millhouse, O. E. (1986) The intercalated cells of the amygdala. Journal of Comparative Neurology 247: 246–71.CrossRefGoogle ScholarPubMed
Minamoto, T. and Kimura, M. (2002) Participation of the thalamic CM-Pf complex in attentional orienting. Journal of Neurophysiology 87: 3090–101.CrossRefGoogle Scholar
Mineff, E. M. and Weinberg, R. J. (2000) Differential synaptic distribution of AMPA receptor subunits in the ventral posterior and reticular thalamic nuclei of the rat. Neuroscience 101: 969–82.CrossRefGoogle ScholarPubMed
Mitchell, S. and Ranck, J. B. (1980) Generation of theta rhythm in medial entorhinal cortex of freely moving rats. Brain Research 189: 49–66.CrossRefGoogle ScholarPubMed
Miyashita, Y. (1993) Inferior temporal cortex: Where visual perception meets memory. Annual Review of Neuroscience 16: 245–63.CrossRefGoogle ScholarPubMed
Mizusawa, K., Ohkoshi, N. and Sasaki, H. (1988) Degeneration of the thalamus and inferior olives associated with spongiform encephalopathy of the cerebral cortex. Clinical Neuropathology 7: 81–6.Google ScholarPubMed
Molinari, M., Minciacchi, D., Bentivoglio, M. and Macchi, G. (1985) Efferent fibers from the motor cortex terminate bilaterally in the thalamus of rats and cats. Experimental Brain Research 57: 305–12.CrossRefGoogle Scholar
Mölle, M., Marshall, L., Gais, S. and Born, J. (2002) Grouping of spindle activity during slow oscillations in human non-REM sleep. Journal of Neuroscience 22: 10941–7.CrossRefGoogle Scholar
Monckton, J. E. and McCormick, D. A. (2002) Neuromodulatory role of serotonin in the ferret thalamus. Journal of Neurophysiology 87: 2124–36.CrossRefGoogle ScholarPubMed
Monckton, J. E. and McCormick, D. A. (2003) Comparative physiological and serotoninergic properties of pulvinar neurons in the monkey, cat and ferret. Thalamus and Related Systems 2: 239–52.Google Scholar
Montero, V. (1986) Localization of gamma-aminobutyric acid (GABA) in type 3 cells and demonstration of their source to F2 terminals in the cat lateral geniculate nucleus. Journal of Comparative Neurology 254: 228–45.CrossRefGoogle ScholarPubMed
Montero, V. (1997) C-FOS induction in sensory pathways of rats exploring a novel complex environment: shifts of active thalamic reticular sectors by predominant sensory cues. Neuroscience 76: 1069–81.CrossRefGoogle ScholarPubMed
Moore, R. Y., Halaris, A. E. and Jones, B. E. (1978) Serotonin neurons of the midbrain raphe: ascending projections. Journal of Comparative Neurology 180: 417–38.CrossRefGoogle ScholarPubMed
Morgan, M. A. and LeDoux, J. E. (1995) Differential contribution of dorsal and ventral medial prefrontal cortex to the acquisition and extinction of conditioned fear in rats. Behavioral Neuroscience 109: 681–8.CrossRefGoogle ScholarPubMed
Morison, R. S. and Bassett, D. L. (1945) Electrical activity of the thalamus and basal ganglia in decorticated cats. Journal of Neurophysiology 8: 309–14.CrossRefGoogle Scholar
Morison, R. S. and Dempsey, E. W. (1942) Mechanism of thalamocortical augmentation and repetition. American Journal of Physiology 138: 297–308.Google Scholar
Morrison, J. H., and Foote, S. L. (1986) Noradrenergic and serotonergic innervation of cortical, thalamic, and tectal visual structures in old and new world monkeys. Journal of Comparative Neurology 243: 117–38.CrossRefGoogle Scholar
Morrison, J. H., Foote, S. L., O'Connor, D. and Bloom, F. E. (1982) Laminar, tangential and regional organization of the noradrenergic innervation of monkey cortex: dopamine-β-hydroxylase immunohistochemistry. Brain Research Bulletin 9: 309–19.CrossRefGoogle ScholarPubMed
Moruzzi, G. (1966) The functional significance of sleep with particular regard to the brain mechanisms underlying consciousness. In Brain and Conscious Experience, ed. Eccles, J. C., pp. 345–79. New York: Springer.Google Scholar
Moruzzi, G. (1972) The sleep-waking cycle. Ergebnisse der Physiologie 64: 1–165.Google ScholarPubMed
Moruzzi, G. and Magoun, H. W. (1949) Brain stem reticular formation and activation of the EEG. Electroencephalography and Clinical Neurophysiology 1: 455–73.CrossRefGoogle ScholarPubMed
Mountcastle, V. B. (1997) The columnar organization of the neocortex. Brain 120: 701–22.CrossRefGoogle ScholarPubMed
Mountcastle, V. B. (1998) Perceptual Neuroscience: The Cerebral Cortex. Cambridge, MA: Harvard University Press.Google Scholar
Mrzljak, L., Pappy, M., Leranth, C. and Goldman-Rakic, P. (1995) Cholinergic synaptic circuitry in the macaque prefrontal cortex. Journal of Comparative Neurology 357: 603–17.CrossRefGoogle ScholarPubMed
Mulle, C., Steriade, M. and Deschênes, M. (1985) The effects of QX-314 on thalamic neurons. Brain Research 333: 350–4.CrossRefGoogle ScholarPubMed
Muller, J., Corodimas, K. P., Fridel, Z. and LeDoux, J. E. (1997) Functional inactivation of the lateral and basal nuclei of the amygdala by muscimol infusion prevents fear conditioning to an explicit conditioned stimulus and to contextual stimuli. Behavioral Neuroscience 111: 683–91.CrossRefGoogle Scholar
Muller, J. F., Mascagni, F. and McDonald, A. J. (2003) Synaptic connections of distinct interneuronal subpopulations in the rat basolateral amygdalar nucleus. Journal of Comparative Neurology 456: 217–36.CrossRefGoogle ScholarPubMed
Mumby, D. G., Wood, E. R. and Pinel, J. (1992) Object-recognition memory is only mildly impaired in rats by lesions of the hippocampus and amygdala. Psychobiology 20: 18–27.Google Scholar
Muñoz, A., Huntsman, M. M. and Jones, E. G. (1998) GAB AB receptor gene expression in monkey thalamus. Journal of Comparative Neurology 394: 118–26.3.0.CO;2-3>CrossRefGoogle Scholar
Murakami, M., Kashiwadani, H., Kirino, Y. and Mori, K. (2005) State-dependent sensory gating in olfactory cortex. Neuron 46: 285–96.CrossRefGoogle ScholarPubMed
Murthy, V. N. and Fetz, E. E. (1992) Coherent 25- to 35-Hz oscillations in the sensorimotor cortex of awake behaving monkeys. Proceedings of the National Academy of Sciences USA 89: 5670–4.CrossRefGoogle ScholarPubMed
Musil, S. Y. and Olson, C. R. (1988) Organization of cortical and subcortical projections to medial prefrontal cortex in the cat. Journal of Comparative Neurology 272: 219–41.CrossRefGoogle ScholarPubMed
Musil, S. Y. and Olson, C. R. (1991) Cortical areas in the medial frontal lobe of the cat delineated by quantitative analysis of thalamic afferents. Journal of Comparative Neurology 308: 457–66.CrossRefGoogle ScholarPubMed
Myers, K. M. and Davis, M. (2002) Behavioral and neural analysis of extinction. Neuron 36: 567–84.CrossRefGoogle ScholarPubMed
Nadasdy, Z., Hirase, H., Czurko, A., Csicsvari, J. and Buzsáki, G. (1999) Replay and time compression of recurring spike sequences in the hippocampus. Journal of Neuroscience 19: 9497–507.CrossRefGoogle ScholarPubMed
Nader, K. (2003) Memory traces unbound. Trends in Neurosciences 26: 65–72.CrossRefGoogle ScholarPubMed
Nader, K., Majidishad, P., Amorapanth, P. and LeDoux, J. E. (2001) Damage to the lateral and central, but not other, amygdaloid nuclei prevents the acquisition of auditory fear conditioning. Learning and Memory 8: 156–63.CrossRefGoogle Scholar
Neckelmann, D., Amzica, F. and Steriade, M. (1998) Spike-wave complexes and fast components of cortically generated seizures. III. Synchronizing mechanisms. Journal of Neurophysiology 80: 1480–94.CrossRefGoogle ScholarPubMed
Nelson, J. P., McCarley, R. W. and Hobson, J. A. (1983) REM sleep burst neurons, PGO waves, and eye movement information. Journal of Neurophysiology 50: 784–97.CrossRefGoogle ScholarPubMed
Neugebauer, V. and Li, W. D. (2003) Differential sensitization of amygdala neurons to afferent inputs. Journal of Neurophysiology 89: 716–27.CrossRefGoogle ScholarPubMed
Nicoll, R. A., Malenka, R. C. and Kauer, J. A. (1990) Functional comparison of neurotransmitter receptor subtypes in mammalian central nervous system. Physiological Reviews 70: 513–65.CrossRefGoogle ScholarPubMed
Niedermeyer, E. (1993) Sleep and EEG. In Electroencephalography: Basic Principles, Clinical Applications and Related Fields (4th edition), ed. Niedermeyer, E. and Silva, F. Lopes da, pp. 153–66. Baltimore, MD: Williams & Wilkins.Google Scholar
Nielsen, T. (2000) Cognition in REM and NREM sleep. Brain and Behavioral Sciences 23: 851–66.CrossRefGoogle ScholarPubMed
Niimi, K., Matsuoka, H., Aisaka, T. and Okada, Y. (1981) Thalamic afferents to the prefrontal cortex in the cat traced with horseradish peroxidase. Journal für Hirnforschung 22: 221–41.Google ScholarPubMed
Nisenbaum, E. S., Xu, Z. C. and Wilson, C. J. (1994) Contribution of a slowly inactivating potassium current to the transition to firing of neostriatal spiny projection neurons. Journal of Neurophysiology 71: 1174–89.CrossRefGoogle ScholarPubMed
Nishimura, Y., Asahi, M., Saitoh, K.et al. (2001) Ionic mechanisms underlying burst firing of layer III sensorimotor cortical neurons of the cat: an in vitro slice study. Journal of Neurophysiology 86: 771–81.CrossRefGoogle Scholar
Nishiyama, M., Hong, K., Mikoshiba, K., Poo, M. M. and Kato, K. (2000) Calcium stores regulate the polarity and input specificity of synaptic modifications. Nature 408: 554–88.Google Scholar
Nita, D., Steriade, M. and Amzica, F. (2003) Hyperpolarisation rectification in cat lateral geniculate neurons modulated by intact corticothalamic projections. Journal of Physiology (London) 552: 325–32.CrossRefGoogle ScholarPubMed
Nitecka, L. and Ben-Ari, Y. (1987) Distribution of GABA-like immunoreactivity in the rat amygdaloid complex. Journal of Comparative Neurology 266: 45–55.CrossRefGoogle ScholarPubMed
Nitecka, L. and Frotscher, M. (1989) Organization and synaptic interconnections of GABAergic and cholinergic elements in the rat amygdaloid nuclei: single- and double-immunolabeling studies. Journal of Comparative Neurology 279: 470–88.CrossRefGoogle ScholarPubMed
Nowak, L. G., Azouz, R., Sanchez-Vives, M. V., Gray, C. M. and McCormick, D. A. (2003) Electrophysiological classes of cat primary visual cortical neurons in vivo as revealed by quantitative analyses. Journal of Neurophysiology 89: 1541–66.CrossRefGoogle ScholarPubMed
Nuñez, A., Amzica, F. and Steriade, M. (1992a) Voltage-dependent fast (20–40 Hz) oscillations in long-axoned neocortical neurons. Neuroscience 51: 7–10.CrossRefGoogle Scholar
Nuñez, A., Amzica, F. and Steriade, M. (1992b) Intrinsic and synaptically generated delta (1–4 Hz) rhythms in dorsal lateral geniculate neurons and their modulation by light-induced fast (30–70 Hz) events. Neuroscience 51: 269–84.CrossRefGoogle Scholar
Nuñez, A., Curró Dossi, R., Contreras, D. and Steriade, M. (1992c) Intracellular evidence for incompatibility between spindle and delta oscillations in thalamocortical neurons of cat. Neuroscience 48: 75–85.CrossRefGoogle Scholar
Nuñez, A., Amzica, F. and Steriade, M. (1993) Electrophysiology of cat association cortical neurons in vivo: intrinsic properties and synaptic responses. Journal of Neurophysiology 70: 418–30.CrossRefGoogle ScholarPubMed
Ojima, H. (1994) Terminal morphology and distribution of corticothalamic fibers originating from layers 5 and 6 of cat primary auditory cortex. Cerebral Cortex 4: 646–63.CrossRefGoogle ScholarPubMed
Öngür, D. and Price, J. L. (2000) The organization of networks within the orbital and medial prefrontal cortex of rats, monkeys and humans. Cerebral Cortex 10: 206–19.CrossRefGoogle ScholarPubMed
Otani, S., Daniel, H., Roisin, M. P. and Crepel, F. (2003) Dopaminergic modulation of long-term synaptic plasticity in rat prefrontal neurons. Cerebral Cortex 13: 1251–6.CrossRefGoogle ScholarPubMed
Packard, M. G. and Teather, L. A. (1998) Amygdala modulation of multiple sensory systems: hippocampus and caudate putamen. Neurobiology of Learning and Memory 69: 163–203.CrossRefGoogle Scholar
Packard, M. G., Cahill, L. and McGaugh, J. L. (1994) Amygdala modulation of hippocampal-dependent and caudate nucleus-dependent memory processes. Proceedings of the National Academy of Sciences USA 91: 8477–81.CrossRefGoogle ScholarPubMed
Pagano, R. R., Gault, M. A. and Gault, F. P. (1964) Amygdala activity: a central measure of arousal. Electroencephalography and Clinical Neurophysiology 17: 255–60.CrossRefGoogle ScholarPubMed
Palva, J. M., Palva, S. and Kaila, K. (2005) Phase synchrony among neuronal oscillations in the human cortex. Journal of Neuroscience 25: 3962–72.CrossRefGoogle ScholarPubMed
Panula, P., Pirvola, U., Auvinen, S. and Airaksinen, M. S. (1989) Histamine-immunoreactive nerve fibers in the rat brain. Neuroscience 28: 585–610.CrossRefGoogle ScholarPubMed
Panula, P., Airaksinen, M. S., Pirvola, U. and Kotilainen, E. (1990) A histamine-containing neuronal system in human brain. Neuroscience 34: 127–32.CrossRefGoogle ScholarPubMed
Pape, H. C. (1996) Queer current and pacemaker: the hyperpolarization-activated cation current in neurons. Annual Reviews of Physiology 58: 299–327.CrossRefGoogle ScholarPubMed
Pape, H. C. and Driesang, R. B. (1998) Ionic mechanisms of intrinsic oscillations in neurons of the basolateral amygdaloid complex. Journal of Neurophysiology 79: 217–26.CrossRefGoogle ScholarPubMed
Pape, H. C. and Mager, R. (1992) Nitric oxide controls oscillatory activity in thalamocortical neurons. Neuron 9: 441–8.CrossRefGoogle ScholarPubMed
Pape, H. C. and McCormick, D. A. (1995) Electrophysiological and pharmacological properties of interneurons in the cat dorsal lateral geniculate nucleus. Neuroscience 68: 1105–25.CrossRefGoogle ScholarPubMed
Pape, H. C., Budde, T., Mager, R. and Kisvárday, Z. F. (1994) Prevention of Ca2+-mediated action potentials in GABAergic local circuit neurones of rat thalamus by a transient K+ current. Journal of Physiology (London) 478: 403–22.CrossRefGoogle Scholar
Pape, H. C., Paré, D. and Driesang, R. B. (1998) Two types of intrinsic oscillations in neurons of the lateral and basolateral nuclei of the amygdala. Journal of Neurophysiology 79: 205–16.CrossRefGoogle ScholarPubMed
Paquet, M. and Smith, Y. (2000) Presynaptic NMDA receptor subunit immunoreactivity in GABAergic terminals in rat brain. Journal of Comparative Neurology 423: 330–47.3.0.CO;2-9>CrossRefGoogle ScholarPubMed
Paré, D. and Collins, D. R. (2000) Neuronal correlates of fear in the lateral amygdala: multiple extracellular recordings in conscious cats. Journal of Neuroscience 20: 2701–10.CrossRefGoogle ScholarPubMed
Paré, D. and Gaudreau, H. (1996) Projection cells and interneurons of the lateral and basolateral amygdala: distinct firing patterns and differential relation to theta and delta rhythms in conscious cats. Journal of Neuroscience 16: 3334–50.CrossRefGoogle ScholarPubMed
Paré, D. and Lang, E. J. (1998) Calcium electrogenesis in neocortical pyramidal neurons in vivo. European Journal of Neuroscience 10: 3164–70.CrossRefGoogle ScholarPubMed
Paré, D. and Llinás, R. (1995) Intracellular study of direct entorhinal inputs to field CA1 in the isolated guinea pig brain in vitro. Hippocampus 5: 115–19.CrossRefGoogle ScholarPubMed
Paré, D. and Smith, Y. (1993a) Distribution of GABA immunoreactivity in the amygdaloid complex of the cat. Neuroscience 57: 1061–76.CrossRefGoogle Scholar
Paré, D. and Smith, Y. (1993b) The intercalated cell masses project to the central and medial nuclei of the amygdala in cats. Neuroscience 57: 1077–90.CrossRefGoogle Scholar
Paré, D. and Smith, Y. (1994) GABAergic projection from the intercalated cell masses of the amygdala to the basal forebrain in cats. Journal of Comparative Neurology 344: 33–49.CrossRefGoogle ScholarPubMed
Paré, D. and Smith, Y. (1996) Thalamic collaterals of corticostriatal axons: their termination field and synaptic targets in cats. Journal of Comparative Neurology 372: 551–67.3.0.CO;2-3>CrossRefGoogle ScholarPubMed
Paré, D. and Steriade, M. (1990) Control of mamillothalamic axis by brainstem cholinergic laterodorsal tegmental afferents: possible involvement in mnemonic processes. In Brain Cholinergic Systems, ed. Steriade, M. and Biesold, D., pp. 337–54. Oxford: Oxford University Press.Google Scholar
Paré, D., Steriade, M., Deschênes, M. and Oakson, G. (1987) Physiological properties of anterior thalamic nuclei, a group devoid of inputs from the reticular thalamic nucleus. Journal of Neurophysiology 57: 1669–85.CrossRefGoogle ScholarPubMed
Paré, D., Smith, Y., Parent, A. and Steriade, M. (1988) Projections of upper brainstem cholinergic and non-cholinergic neurons of cat to intralaminar and reticular thalamic nuclei. Neuroscience 25: 69–88.CrossRefGoogle ScholarPubMed
Paré, D., Smith, Y., Parent, A. and Steriade, M. (1989) Neuronal activity of identified posterior hypothalamic neurons projecting to the brainstem peribrachial area of the cat. Neuroscience Letters 107: 145–50.CrossRefGoogle ScholarPubMed
Paré, D., Curró Dossi, R., Datta, S. and Steriade, M. (1990a) Brainstem genesis of reserpine-induced ponto-geniculo-occipital waves: an electrophysiological and morphological investigation. Experimental Brain Research 81: 533–44.CrossRefGoogle Scholar
Paré, D., Steriade, M., Deschênes, M. and Bouhassira, D. (1990b) Prolonged enhancement of anterior thalamic synaptic responsiveness by stimulation of a brainstem cholinergic group. Journal of Neuroscience 10: 20–33.CrossRefGoogle Scholar
Paré, D., Curró Dossi, R. and Steriade, M. (1991) Three types of inhibitory postsynaptic potentials generated by interneurons in the anterior thalamic complex of cat. Journal of Neurophysiology 66: 1190–204.CrossRefGoogle ScholarPubMed
Paré, D., Dong., J. and Gaudreau, H. (1995a) Amygdalo-entorhinal relations and their reflection in the hippocampal formation: generation of sharp sleep potentials. Journal of Neuroscience 15: 2482–503.CrossRefGoogle Scholar
Paré, D., Pape, H. C. and Dong, J. (1995b) Bursting and oscillating neurons of the cat basolateral amygdaloid complex in vivo: electrophysiological properties and morphological features. Journal of Neurophysiology 74: 1179–91.CrossRefGoogle Scholar
Paré, D., Smith, Y. and Paré, J. F. (1995c) Intra-amygdaloid projections of the basolateral and basomedial nuclei in the cat: Phaseolus vulgaris-leucoagglutinin anterograde tracing at the light and electron microscopic level. Neuroscience 69: 567–83.CrossRefGoogle Scholar
Paré, D., Shink, E., Gaudreau, H., Destexhe, A. and Lang, E. J. (1998a) Impact of spontaneous synaptic activity on the resting properties of cat neocortical pyramidal neurons in vivo. Journal of Neurophysiology 79: 1450–60.CrossRefGoogle Scholar
Paré, D., Lang, E. J. and Destexhe, A. (1998b) Inhibitory control of somatodendritic interactions underlying action potentials in neocortical pyramidal neurons in vivo: an intracellular and computational study. Neuroscience 84: 377–402.CrossRefGoogle Scholar
Paré, D., Collins, D. R. and Pelletier, J. G. (2002).Amygdala oscillations and the consolidation of emotional memories. Trends in Cognitive Sciences 6: 306–14.CrossRefGoogle ScholarPubMed
Paré, D., Quirk, G. J. and LeDoux, J. E. (2004) New vistas on amygdala networks in conditioned fear. Journal of Neurophysiology 92: 1–9.CrossRefGoogle ScholarPubMed
Parent, A., Descarries, L. and Beaudet, A. (1981) Organization of ascending serotonin systems in the adult rat brain. A radioautographic study after intraventricular administration of (3H)5-hydroxytryptamine. Neuroscience 6: 115–38.CrossRefGoogle Scholar
Parent, A., Paré, D., Smith, Y., and Steriade, M. (1988) Basal forebrain cholinergic and non-cholinergic projections to the thalamus and brainstem in cats and monkeys. Journal of Comparative Neurology 277: 281–301.CrossRefGoogle Scholar
Parent, M. B., Quirarte, G. L., Cahill, L. and McGaugh, J. L. (1995) Spared retention of inhibitory avoidance learning after postraining amygdala lesions. Behavioral Neuroscience 109: 803–7.CrossRefGoogle Scholar
Parmentier, R., Ohtsu, H., Djebbara-Hannas, Z.et al. (2002) Anatomical, physiological, and pharmacological characteristics of histidine decarboxylase knock-out mice: evidence for the role of brain histamine in behavioral and sleep-wake control. Journal of Neuroscience 22: 7695–711.CrossRefGoogle ScholarPubMed
Pavlides, C. and Winson, J. (1989) Influences of hippocampal place cell firing in the awake state on the activity of these cells during subsequent sleep episodes. Journal of Neuroscience 9: 2907–18.CrossRefGoogle ScholarPubMed
Pavlides, C., Greenstein, Y. J., Goudman, M. and Winson, J. (1988) Long-term potentiation in the dentate gyrus is induced preferentially on the positive phase of the theta-rhythm. Brain Research 439: 383–7.CrossRefGoogle ScholarPubMed
Pavlov, I. P. (1923) ‘Innere Hemmung’ der bedingten Reflexe und der Schlaf – ein und derselbe Prozess. Skandinavische Archive für Physiologie 44: 42–58.Google Scholar
Pavlov, I. P. (1927) Conditioned Reflexes: An Investigation of the Physiological Activity of the Cerebral Cortex. London: Oxford University Press.
Paxinos, G. and Watson, C. (1986) The Rat Brain in Stereotaxic Coordinates. New York: Academic Press.Google Scholar
Pedroarena, C. and Llinás, R. (1997) Dendritic calcium conductances generate high-frequency oscillation in thalamocortical neurons. Proceedings of the National Academy of Sciences USA 94: 724–8.CrossRefGoogle ScholarPubMed
Pelletier, J. G. and Paré, D. (2002) Uniform range of conduction times from the lateral amygdala to distributed perirhinal sites. Journal of Neurophysiology 87: 1213–21.CrossRefGoogle ScholarPubMed
Pelletier, J. G., Apergis, J. and Paré, D. (2004) Low probability transmission of neocortical and entorhinal impulses through the perirhinal cortex. Journal of Neurophysiology 91: 2079–89.CrossRefGoogle ScholarPubMed
Pelletier, J. G., Apergis-Schoute, J. and Paré, D. (2005a) Interaction between amygdala and neocortical inputs in the perirhinal cortex. Journal of Neurophysiology 94: 1837–48.CrossRefGoogle Scholar
Pelletier, J. G., Likhtik, E., Filali, M. and Paré, D. (2005b) Lasting increases in basolateral amygdala activity after emotional arousal: implications for facilitated consolidation of emotional memories. Learning and Memory 12: 96–102.CrossRefGoogle Scholar
Peña, E. and Geijo-Barrientos, E. (1996) Laminar localization, morphology, and physiological properties of pyramidal neurons that have low-threshold calcium current in the guinea-pig medial frontal cortex. Journal of Neuroscience 16: 5301–11.CrossRefGoogle ScholarPubMed
Pennartz, C. M., Uylings, H. B., Barnes, C. A. and McNaughton, B. L. (2002) Memory reactivation and consolidation during sleep: from cellular mechanisms to human performance. Progress in Brain Research 138: 143–66.CrossRefGoogle ScholarPubMed
Pennartz, C. M., Lee, E., Verheul, J.et al. (2004) The ventral striatum in off-line processing: ensemble reactivation during sleep and modulation by hippocampal ripples. Journal of Neuroscience 24: 6446–56.CrossRefGoogle ScholarPubMed
Perez-Velazquez, J. L. and Carlen, P. L. (2000) Gap junctions, synchrony and seizures. Trends in Neurosciences 23: 68–74.CrossRefGoogle ScholarPubMed
Perreault, M. C., Qin, Y., Heggelund, P. and Zhu, J. J. (2003) Postnatal development of GABAergic signalling in the rat lateral geniculate nucleus: presynaptic dendritic mechanisms. Journal of Physiology (London) 546: 137–48.CrossRefGoogle ScholarPubMed
Petersen, C. C. H., Hahn, T. T. G., Mehta, M., Grinvald, A. and Sakmann, B. (2003) Interaction of sensory responses with spontaneous depolarization in layer 2/3 barrel cortex. Proceedings of the National Academy of Sciences USA 100: 13638–43.CrossRefGoogle ScholarPubMed
Petrides, M. and Pandya, D. N. (1984) Projections to frontal cortex from posterior parietal region in rhesus monkeys. Journal of Comparative Neurology 228: 105–16.CrossRefGoogle Scholar
Petrovich, G. D. and Swanson, L. W. (1997) Projections from the lateral part of the central amygdalar nucleus to the postulated fear conditioning circuit. Brain Research 763: 247–54.CrossRefGoogle ScholarPubMed
Petsche, H., Stumpf, C. and Gogolak, G. (1962) The significance of the rabbit's septum as a relay station between the midbrain and the hippocampus. I. The control of hippocampus arousal activity by the septum cells. Electroencephalography and Clinical Neurophysiology 14: 202–11.CrossRefGoogle ScholarPubMed
Petsche, H., Pockberger, H. and Rappelsberger, P. (1984) On the search for the sources of the electroencephalogram. Neuroscience 11: 1–27.CrossRefGoogle ScholarPubMed
Peyron, C., Tighe, D. K., Pol, A. N.et al. (1998) Neurons containing hypocretin (orexin) project to multiple neuronal systems. Journal of Neuroscience 18: 9996–10015.CrossRefGoogle ScholarPubMed
Phelps, E. A. (2004) Human emotion and memory: interactions of the amygdala and hippocampal complex. Current Opinion in Neurobiology 14: 198–202.CrossRefGoogle ScholarPubMed
Phelps, E. A., Delgado, M. R., Nearing, K. I. and LeDoux, J. E. (2004) Extinction learning in humans: role of the amygdala and vmPFC. Neuron 43: 897–905.CrossRefGoogle ScholarPubMed
Pickel, V. M., Joh, T. H. and Reis, D. J. (1977) A serotonergic innervation of noradrenergic neurons in nucleus locus coeruleus: demonstration by immunocytochemical localization of the transmitter specific enzymes tyrosine and tryptophan hydroxylase. Brain Research 131: 197–214.CrossRefGoogle ScholarPubMed
Pickel, V. M., Van, B. E., Chan, J. and Cestari, D. M. (1995) Amygdala efferents form inhibitory-type synapses with a subpopulation of catecholaminergic neurons in the rat nucleus tractus solitarius. Journal of Comparative Neurology 362: 510–23.CrossRefGoogle ScholarPubMed
Pickel, V. M., Van, B. E., Chan, J. and Cestari, D. M. (1996) GABAergic neurons in rat nuclei of solitary tracts receive inhibitory-type synapses from amygdaloid efferents lacking detectable GABA-immunoreactivity. Journal of Neuroscience Research 44: 446–58.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Pikkarainen, M., Rönkkö, S., Savander, V., Insausti, R. and Pitkänen, A. (1999) Projections from the lateral, basal, and accessory basal nuclei of the amygdala to the hippocampal formation in rat. Journal of Comparative Neurology 403: 229–60.3.0.CO;2-P>CrossRefGoogle ScholarPubMed
Pinault, D., Smith, Y. and Deschênes, M. (1997) Dendrodendritic and axoaxonic synapses in the thalamic reticular nucleus of the adult rat. Journal of Neuroscience 17: 3215–33.CrossRefGoogle ScholarPubMed
Pinault, D., Leresche, N., Charpier, S.et al. (1998) Intracellular recordings in thalamic neurones during spontaneous spike and wave discharges in rats with absence epilepsy. Journal of Physiology (London) 509: 449–56.CrossRefGoogle ScholarPubMed
Pinto, A. and Sesack, S. R. (2002) Prefrontal cortex projection to the rat amygdala: ultrastructural relationship to dopamine D1 and D2 receptors. Society for Neuroscience Abstracts 28: 587.Google Scholar
Pinto, A., Fuentes, C. and Paré, D. (2006) Feedforward inhibition regulates perirhinal transmission of neocortical inputs to the entorhinal cortex: Ultrastructural study in guinea pigs. Journal of Comparative Neurology 495: 722–34.CrossRefGoogle ScholarPubMed
Pitkänen, A. (2000) Connectivity of the rat amygdaloid complex. In The Amygdala: A Functional Analysis, ed. Aggleton, J. P., pp. 31–115. Oxford: Oxford University Press.Google Scholar
Pitkänen, A. and Amaral, D. G. (1993a) Distribution of calbindin-D28k immunoreactivity in the monkey temporal lobe: the amygdaloid complex. Journal of Comparative Neurology 331: 199–224.CrossRefGoogle Scholar
Pitkänen, A. and Amaral, D. G. (1993b) Distribution of parvalbumin-immunoreactive cells and fibers in the monkey temporal lobe: the amygdaloid complex. Journal of Comparative Neurology 331: 14–36.CrossRefGoogle Scholar
Pitkänen, A. and Amaral, D. G. (1994) The distribution of GABAergic cells, fibers, and terminals in the monkey amygdaloid complex: an immunohistochemical and in situ hybridization study. Journal of Neuroscience 14: 2200–24.CrossRefGoogle ScholarPubMed
Pitkänen, A., Stefanacci, L., Farb, C. R.et al. (1995) Intrinsic connections of the rat amygdaloid complex: projections originating in the lateral nucleus. Journal of Comparative Neurology 356: 288–310.CrossRefGoogle ScholarPubMed
Pitkänen, A., Savander, V. and LeDoux, J. E. (1997) Organization of intra-amygdaloid circuitries in the rat: an emerging framework for understanding functions of the amygdala. Trends in Neurosciences 20: 517–23.CrossRefGoogle ScholarPubMed
Pitkänen, A., Pikkarainen, M., Nurminen, N. and Ylinen, A. (2000) Reciprocal connections between the amygdala and the hippocampal formation, perirhinal cortex, and postrhinal cortex in rat. Annals of the New York Academy of Sciences 911: 369–91.CrossRefGoogle ScholarPubMed
Pitkänen, A., Kelly, J. L. and Amaral, D. G. (2002) Projections from the lateral, basal, and accessory basal nuclei of the amygdala to the entorhinal cortex in the macaque monkey. Hippocampus 12: 186–205.CrossRefGoogle ScholarPubMed
Pivik, T. and Foulkes, D. (1968) NREM mentation: relation to personality, orientation time and time of night. Journal of Consultation Clinical Psychology 32: 144–51.CrossRefGoogle ScholarPubMed
Plihal, W. and Born, J. (1997) Effects of early and late nocturnal sleep on declarative and procedural memory. Journal of Cognitive Neuroscience 9: 534–47.CrossRefGoogle ScholarPubMed
Plum, F. (1991) Coma and related global disturbances of the human conscious state. In Cerebral Cortex, vol. 9 (Normal and Altered States of Function), ed. Peters, A. and Jones, E. G., pp. 359–425. New York: Plenum.CrossRefGoogle Scholar
Poggio, G. F. and Mouncastle, V. B. (1960) A study in the functional contributions of the lemniscal and spinothalamic systems to somatic sensibility: central nervous mechanisms in pain. Bulletin of the Johns Hopkins Hospital 106: 266–316.Google ScholarPubMed
Popescu, A. T., Saghyan, A. and Paré, D. (2005) Amygdala activation opens temporal windows of plasticity in the cortico-striatal pathway. Society for Neuroscience Abstracts 31: 651–4.Google Scholar
Power, A. E. (2004) Slow-wave sleep, acetylcholine, and memory consolidation. Proceedings of the National Academy of Sciences USA 101: 1795–6.CrossRefGoogle ScholarPubMed
Preuss, T. M. and Goldman-Rakic, P. S. (1987) Crossed corticothalamic and thalamocortical connections of macaque prefrontal cortex. Journal of Comparative Neurology 257: 269–81.CrossRefGoogle ScholarPubMed
Pritzel, M. and Markowitsch, H. J. (1981) Cortico-prefrontal afferents in the guinea pig. Brain Research Bulletin 7: 427–34.CrossRefGoogle ScholarPubMed
Puil, E., Meiri, H. and Yarom, Y. (1994) Resonant behavior and frequency preferences of thalamic neurons. Journal of Neurophysiology 71: 575–82.CrossRefGoogle ScholarPubMed
Purpura, D. P. (1970) Operations and processes in thalamic and synaptically related neural subsystems. In The Neuroscience: Second Study Program, ed. Schmitt, F. O., pp. 458–470. New York: Rockefeller University Press.Google Scholar
Purpura, D. P., McMurtry, J. G. and Maekawa, K. (1966) Synaptic events in ventrolateral thalamic neurons during suppression of recruiting responses by brain stem reticular stimulation. Brain Research 1: 63–76.CrossRefGoogle ScholarPubMed
Quirk, G. J., Repa, J. C. and LeDoux, J. E. (1995) Fear conditioning enhances short-latency auditory responses of lateral amygdala neurons: Parallel recordings in the freely behaving rat. Neuron 15: 1029–39.CrossRefGoogle ScholarPubMed
Quirk, G. J., Russo, G. K., Barron, J. L. and Lebron, K. (2000) The role of ventromedial prefrontal cortex in the recovery of extinguished fear. Journal of Neuroscience 20: 6225–31.CrossRefGoogle ScholarPubMed
Quirk, G. J., Likhtik, E., Pelletier, J. G. and Paré, D. (2003) Stimulation of medial prefrontal cortex decreases the responsiveness of central amygdala output neurons. Journal of Neuroscience 23: 8800–7.CrossRefGoogle ScholarPubMed
Raghavachari, S., Kahana, M. J., Rizzuto, D. S.et al. (2001) Gating of human theta oscillations by a working memory task. Journal of Neuroscience 21: 3175–83.CrossRefGoogle ScholarPubMed
Rainnie, D. G., Asprodini, E. K. and Shinnick-Gallagher, G. P. (1991) Inhibitory transmission in the basolateral amygdala. Journal of Neurophysiology 66: 999–1009.CrossRefGoogle ScholarPubMed
Rainnie, D. G., Asprodini, E. K. and Shinnick-Gallagher, G. P. (1993) Intracellular recordings from morphologically identified neurons of the basolateral amygdala. Journal of Neurophysiology 69: 1350–62.CrossRefGoogle ScholarPubMed
Ramm, P. and Smith, C. T. (1990) Rates of cerebral protein synthesis are linked to slow wave sleep in the rat. Physiology and Behavior 48: 749–53.CrossRefGoogle ScholarPubMed
Rao, Z. R., Shiosaka, S. and Tohyama, M. (1987) Origin of cholinergic fibers in the basolateral nucleus of the amygdaloid complex using sensitive double-labeling technique of retrograde biotinized tracer and immunocytochemistry. Journal für Hirnforschung 28: 553–60.Google ScholarPubMed
Rash, J. E., Staines, W. A., Yasumura, T.et al. (2000) Immunogold evidence that neuronal gap junctions in adult rat brain and spinal cord contain connexin-36 but not connexin-32 or connexin-43. Proceedings of the National Academy of Sciences USA 97: 7573–8.CrossRefGoogle ScholarPubMed
Rasler, F. E. (1984) Behavioral and electrophysiological manifestations of bombesin: excessive grooming and elimination of sleep. Brain Research 321: 187–98.CrossRefGoogle Scholar
Rasmusson, D. D., Clow, K. and Szerb, J. C. (1994) Modification of neocortical acetylcholine release and electroencephalogram desynchronization due to brainstem stimulation by drugs applied to the basal forebrain. Neuroscience 60: 665–77.CrossRefGoogle ScholarPubMed
Rasmusson, D. D., Szerb, J. C. and Jordan, J. L. (1996) Differential effects of α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid and N-methyl-D-aspartate receptor antagonists applied to the basal forebrain on cortical acetylcholine release and EEG desynchronization. Neuroscience 72: 419–27.CrossRefGoogle Scholar
Rechtschaffen, A. (1998) Current perspective on the function of sleep. Perspectives in Biology and Medicine 41: 359–90.CrossRefGoogle Scholar
Rechtschaffen, A., Verdone, P. and Wheaton, J. (1963) Reports of mental activity during sleep. Canadian Journal of Psychiatry 8: 409–14.Google ScholarPubMed
Reinagel, P., Godwin, D., Sherman, S. M. and Koch, C. (1999) Encoding of visual information by LGN bursts. Journal of Neurophysiology 81: 2558–69.CrossRefGoogle ScholarPubMed
Rempel-Clower, N. L. and Barbas, H. (1998) Topographic organization of connections between the hypothalamus and prefrontal cortex in the rhesus monkey. Journal of Comparative Neurology 398: 393–419.3.0.CO;2-V>CrossRefGoogle ScholarPubMed
Repa, J. C., Muller, J., Apergis, J.et al. (2001) Two different lateral amygdala cell populations contribute to the initiation and storage of memory. Nature Neuroscience 4: 724–31.CrossRefGoogle Scholar
Rescorla, R. A. (2004) Spontaneous recovery. Learning and Memory 11: 501–9.CrossRefGoogle ScholarPubMed
Ressler, K. J., Rothbaum, B. O., Tannenbaum, L.et al. (2004) Cognitive enhancers as adjuncts to psychotherapy: use of D-cycloserine in phobic individuals to facilitate extinction of fear. Archives of General Psychiatry 61: 1136–44.CrossRefGoogle Scholar
Ribary, U., Ioannides, A. A., Singh, K. D.et al. (1991) Magnetic field tomography of coherent thalamocortical 40-Hz oscillations in humans. Proceedings of the National Academy of Sciences USA 88: 11037–41.CrossRefGoogle ScholarPubMed
Rice, D. P. and Miller, L. S. (1995) The economic burden of affective disorders. British Journal of Psychiatry Suppl. 27: 34–42.Google Scholar
Richardson, M. P., Strange, B. A. and Dolan, R. J. (2004) Encoding of emotional memories depends on amygdala and hippocampus and their interactions. Nature Neuroscience 7: 278–85.CrossRefGoogle ScholarPubMed
Riches, I. P., Wilson, F. A. and Brown, M. W. (1991) The effects of visual stimulation and memory on neurons of the hippocampal formation and the neighboring parahippocampal gyrus and inferior temporal cortex of the primate. Journal of Neuroscience 11: 1763–79.CrossRefGoogle ScholarPubMed
Rinzel, J., Terman, D., Wang, X. J. and Ermentrout, B. (1998) Propagating activity patterns in large-scale inhibitory neuronal networks. Science 279: 1351–5.CrossRefGoogle ScholarPubMed
Robbins, T. W. and Everitt, B. J. (1995) Arousal systems and attention. In The Cognitive Neurosciences, ed. Gazzaniga, M. S., pp. 703–20. Cambridge, MA: MIT Press.
Robinson, T. E. (1980) Hippocampal rhythmic slow activity (RSA; theta): a critical analysis of selected studies and discussion of possible species-differences. Brain Research 203: 69–101.CrossRefGoogle ScholarPubMed
Roffwarg, H. P., Muzio, J. N. and Dement, W. C. (1966) Ontogenetic development of the human sleep-dream cycle. Science 152: 604–19.CrossRefGoogle ScholarPubMed
Rogan, M. T. and LeDoux, J. E. (1995) LTP is accompanied by commensurate enhancement of auditory-evoked responses in a fear conditioning circuit. Neuron 15: 127–36.CrossRefGoogle Scholar
Rogan, M. T., Stäubli, U. V. and LeDoux, J. E. (1997) Fear conditioning induces associative long-term potentiation in the amygdala. Nature 390: 604–7.CrossRefGoogle ScholarPubMed
Rolls, E. T., Miyashita, Y., Cahusac, P. M. B.et al. (1989) Hippocampal neurons in the monkey with activity related to the place in which a stimulus is shown. Journal of Neuroscience 9: 1835–45.CrossRefGoogle Scholar
Rolls, E. T., Cahusac, P., Feigenbaum, J. D. and Miyashita, Y. (1993) Responses of single neurons in the hippocampus of the macaque related to recognition memory. Experimental Brain Research 93: 299–306.CrossRefGoogle ScholarPubMed
Rolls, E. T., Inoue, K. and Browning, A. (2003) Activity of primate subgenual cingulate cortex neurons is related to sleep. Journal of Neurophysiology 90: 134–42.CrossRefGoogle ScholarPubMed
Romanski, L. M., Clugnet, M. C., Bordi, F. and LeDoux, J. E. (1993) Somatosensory and auditory convergence in the lateral nucleus of the amygdala. Behavioral Neuroscience 107: 444–50.CrossRefGoogle ScholarPubMed
Room, P. and Groenewegen, H. J. (1986a) Connections of the parahippocampal cortex. I. Cortical afferents. Journal of Comparative Neurology 251: 415–50.CrossRefGoogle Scholar
Room, P. and Groenewegen, H. J. (1986b) Connections of the parahippocampal cortex in the cat. II. Subcortical afferents. Journal of Comparative Neurology 251: 451–73.CrossRefGoogle Scholar
Room, P., Russchen, F. T., Groenewegen, H. J. and Lohman, A. H. (1985) Efferent connections of the prelimbic (area 32) and the infralimbic (area 25) cortices: an anterograde tracing study in the cat. Journal of Comparative Neurology 242: 40–55.CrossRefGoogle ScholarPubMed
Roozendaal, B. (2000) Glucocorticoids and the regulation of memory consolidation. Psychoneuroendocrinology 25: 213–38.CrossRefGoogle ScholarPubMed
Roozendaal, B. and McGaugh, J. L. (1996a) Amygdaloid nuclei lesions differentially affect glucocorticoid-induced memory enhancement in an inhibitory avoidance task. Neurobiology of Learning and Memory 65: 1–8.CrossRefGoogle Scholar
Roozendaal, B. and McGaugh, J. L. (1996b) The memory-modulatory effects of glucocorticoids depend on an intact stria terminalis. Brain Research 709: 243–50.CrossRefGoogle Scholar
Ropert, N. and Steriade, M. (1981) Input-output organization of the midbrain reticular core. Journal of Neurophysiology 46: 17–31.CrossRefGoogle ScholarPubMed
Rosanova, M. and Timofeev, I. (2005) Neuronal mechanisms mediating the variability of somatosensory evoked potentials during sleep oscillations in cats. Journal of Physiology (London) 562: 569–82.CrossRefGoogle ScholarPubMed
Rose, J. E. and Woolsey, C. N. (1948) The orbitofrontal cortex and its connections with the mediodorsal nucleus in rabbit, sheep and cat. Publications on Research of Nervous and Mental Disease 27: 210–32.Google Scholar
Rosen, J. B., Hitchcock, J. M., Sananes, C. B., Miserendino, M. and Davis, M. (1991) A direct projection from the central nucleus of the amygdala to the acoustic startle pathway: anterograde and retrograde tracing studies. Behavioral Neuroscience 105: 817–25.CrossRefGoogle ScholarPubMed
Rosene, D. L. and Va, Hoesen, G. W. (1977) Hippocampal efferents reach widespread areas of cerebral cortex and amygdala in the rhesus monkey. Science 198: 315–17.CrossRefGoogle ScholarPubMed
Rosenkranz, J. A. and Grace, A. A. (1999) Modulation of basolateral amygdala neuronal firing and afferent drive by dopamine receptor activation in vivo. Journal of Neuroscience 19: 11027–39.CrossRefGoogle ScholarPubMed
Rosenkranz, J. A. and Grace, A. A. (2001) Dopamine attenuates prefrontal cortical suppression of sensory inputs to the basolateral amygdala of rats. Journal of Neuroscience 21: 4090–103.CrossRefGoogle ScholarPubMed
Rosenkranz, J. A. and Grace, A. A. (2002) Dopamine-mediated modulation of odour-evoked amygdala potentials during Pavlovian conditioning. Nature 417: 282–7.CrossRefGoogle ScholarPubMed
Rossi, G., Macchi, G., Porro, M.et al. (1998) Fatal familial insomnia. Genetic, neuropathologic, and biochemical study of a patient from a new Italian kindred. Neurology 50: 688–92.CrossRefGoogle ScholarPubMed
Roth, M., Shaw, J. and Green, J. (1956) The form, voltage distribution and physiological significance of the K-complex. Electroencephalography and Clinical Neurophysiology 8: 385–402.CrossRefGoogle ScholarPubMed
Rothbaum, B. O. and Davis, M. (2003) Applying learning principles to the treatment of post-trauma reactions. Annals of the New York Academy of Sciences 1008: 112–21.CrossRefGoogle ScholarPubMed
Rougeul-Buser, A., Bouyer, J. J., Montaron, M. F. and Buser, P. (1983) Patterns of activities in the ventrobasal thalamus and somatic cortex SI during behavioural immobility in the awake cat: focal waking rhythms. Experimental Brain Research Suppl. 7: 69–87.CrossRefGoogle Scholar
Rowell, P. P., Volk, K. A., Li, J. and Bickford, M. E. (2003) Investigations of the cholinergic modulation of GABA release in rat thalamic slices. Neuroscience 116: 447–53.CrossRefGoogle Scholar
Roy, J. P., Clercq, M., Steriade, M. and Deschênes, M. (1984) Electrophysiology of neurons of the lateral thalamic nuclei in cat: mechanisms of long-lasting hyperpolarizations. Journal of Neurophysiology 51: 1220–35.CrossRefGoogle ScholarPubMed
Royer, S. and Paré, D. (2002) Bidirectional synaptic plasticity in intercalated amygdala neurons and the extinction of conditioned fear responses. Neuroscience 115: 455–62.CrossRefGoogle ScholarPubMed
Royer, S. and Paré, D. (2003) Conservation of total synaptic weights via inverse homo- vs. heterosynaptic LTD and LTP. Nature 422: 518–22.CrossRefGoogle Scholar
Royer, S., Martina, M. and Paré, D. (1999) An inhibitory interface gates impulse traffic between the input and output stations of the amygdala. Journal of Neuroscience 19: 10575–83.CrossRefGoogle ScholarPubMed
Royer, S., Martina, M. and Paré, D. (2000a) Bistable behavior of inhibitory neurons controlling impulse traffic through the amygdala: role of a slowly deinactivating K+ current. Journal of Neuroscience 20: 9034–9.CrossRefGoogle Scholar
Royer, S., Martina, M. and Paré, D. (2000b) Polarized synaptic interactions between intercalated neurons of the amygdala. Journal of Neurophysiology 83: 3509–18.CrossRefGoogle Scholar
Ruch-Monachon, M. A., Jalfre, M. and Haefeley, W. (1976) Drugs and PGO waves in the lateral geniculate body of the curarized rat. Archives Internationales de Pharmacodynamie et Thérapie 219: 251–346.Google Scholar
Rudolph, M. and Destexhe, A. (2003) The discharge variability of neocortical neurons during high-conductance states. Neuroscience 119: 855–73.CrossRefGoogle ScholarPubMed
Rudy, B. and McBain, C. J. (2001) Kv3 channels: voltage-gated K+ channels designed for high-frequency repetitive firing. Trends in Neurosciences 24: 517–26.CrossRefGoogle ScholarPubMed
Russchen, F. T. (1982a) Amygdalopetal projections in the cat. I. Cortical afferent connections. A study with retrograde and anterograde tracing techniques. Journal of Comparative Neurology 206: 159–79.CrossRefGoogle Scholar
Russchen, F. T. (1982b) Amygdalopetal projections in the cat. II. Subcortical afferent connections. A study with retrograde tracing techniques. Journal of Comparative Neurology 207: 157–76.CrossRefGoogle Scholar
Ruth, R. E., Collier, T. J. and Routtenberg, A. (1982) Topography between the entorhinal cortex and the dentate septotemporal axis in rats: I. Medial and intermediate entorhinal projecting cells. Journal of Comparative Neurology 209: 69–78.CrossRefGoogle ScholarPubMed
Rye, D. B., Saper, C. B., Lee, H. J. and Wainer, B. H. (1987) Pedunculopontine tegmental nucleus of the rat: cytoarchitecture, cytochemistry, and some extrapyramidal connections of the mesopontine tegmentum. Journal of Comparative Neurology 259: 483–528.CrossRefGoogle ScholarPubMed
Sacchetti, B., Lorenzini, C. A., Baldi, E., Tassoni, G. and Bucherelli, C. (1999) Auditory thalamus, dorsal hippocampus, basolateral amygdala, and perirhinal cortex role in the consolidation of conditioned freezing to context and to acoustic conditioned stimulus in the rat. Journal of Neuroscience 19: 9570–8.CrossRefGoogle ScholarPubMed
Sainsbury, R. S. (1998) Hippocampal theta: a sensory-inhibition theory of function. Neuroscience and Biobehavioral Reviews 22: 237–41.CrossRefGoogle ScholarPubMed
Sakai, K. (1985) Anatomical and physiological basis of paradoxical sleep. In Brain Mechanisms of Sleep, ed. McGinty, D. J., Morrison, A., Drucker-Colin, R. and Parmeggiani, P. L., pp. 111–37. New York: Raven.Google Scholar
Sakai, K. and Crochet, S. (2001) Differentiation of presumed serotonergic dorsal raphe neurons in relation to behavior and wake-sleep states. Neuroscience 104: 1141–55.CrossRefGoogle ScholarPubMed
Sakai, K. and Jouvet, M. (1980) Brainstem PGO-on cells projecting directly to the cat lateral geniculate nucleus. Brain Research 194: 500–5.CrossRefGoogle Scholar
Sakai, K., Salvert, D., Touret, M. and Jouvet, M. (1977a) Afferent connections of the nucleus raphe dorsalis in the cat as visualized by the horseradish peroxidase technique. Brain Research 137: 11–35.CrossRefGoogle Scholar
Sakai, K., Touret, M., Salvert, D., Leger, L. and Jouvet, M. (1977b) Afferent projections to the cat locus coeruleus as visualized by the horseradish peroxidase technique. Brain Research 119: 21–41.CrossRefGoogle Scholar
Sakai, K., El Mansari, M., Lin, J. S., Zhang, G. and Vanni-Mercier, F. (1990) The posterior hypothalamus in the regulation of wakefulness and paradoxical sleep. In The Diencephalon and Sleep, ed. Mancia, M. and Marini, G., pp. 171–98. New York: Raven.Google Scholar
Sakakura, H. (1968) Spontaneous and evoked unitary activities of cat lateral geniculate neurons in sleep and wakefulness. Japanese Journal of Physiology 18: 23–42.CrossRefGoogle ScholarPubMed
Salami, M., Itami, C., Tsumoto, T. and Kimura, F. (2003).Change in conduction velocity by regional myelination yields constant latency irrespective of distance between thalamus and cortex. Proceedings of the National Academy of Sciences USA 100: 6174–9.CrossRefGoogle ScholarPubMed
Salinas, J. A., Introini-Collison, I. B., Dalmaz, C. and McGaugh, J. L. (1997) Posttraining intraamygdala infusions of oxotremorine and propranolol modulate storage of memory for reductions in reward magnitude. Neurobiology of Learning and Memory 68: 51–9.CrossRefGoogle ScholarPubMed
Sallanon, M., Denoyer, M., Kitahama, K., Aubert, C., Gay, N. and Jouvet, M. (1989) Long-lasting insomnia induced by preoptic lesions and its transient reversal by muscimol injection into the posterior hypothalamus in the cat. Neuroscience 32: 669–83.CrossRefGoogle ScholarPubMed
Salt, T. E. (1989) Gamma-aminobutyric acid and afferent inhibition in the cat and rat ventrobasal thalamus. Neuroscience 28: 17–26.CrossRefGoogle ScholarPubMed
Samson, R. D. and Paré, D. (2005) Activity-dependent synaptic plasticity in the central nucleus of the amygdala. Journal of Neuroscience 25: 1847–55.CrossRefGoogle ScholarPubMed
Sanchez, R. and Leonard, C. S. (1996) NMDA-receptor-mediated synaptic currents in guinea pig laterodorsal tegmental neurons in vitro. Journal of Neurophysiology 76: 1101–11.CrossRefGoogle ScholarPubMed
Sanchez-Vives, M. V. and McCormick, D. A. (1997) Functional properties of perigeniculate inhibition of dorsal lateral geniculate nucleus thalamocortical neurons in vitro. Journal of Neuroscience 17: 8880–93.CrossRefGoogle ScholarPubMed
Sanchez-Vives, M. V. and McCormick, D. A. (2000) Cellular and network mechanisms of rhythmic recurrent activity in neocortex. Nature Neuroscience 3: 1027–34.CrossRefGoogle ScholarPubMed
Sanchez-Vives, M. V., Bal, T. and McCormick, D. A. (1997) Inhibitory interactions between perigeniculate GABAergic neurons. Journal of Neuroscience 17: 8894–908.CrossRefGoogle ScholarPubMed
Sanders, M. J., Wiltgen, B. J. and Fanselow, M. S. (2003) The place of the hippocampus in fear conditioning. European Journal of Pharmacology 463: 217–23.CrossRefGoogle ScholarPubMed
Santini, E., Ge, H., Ren, K., Pena de Ortiz, S. and Quirk, G. J. (2004) Consolidation of fear extinction requires protein synthesis in the medial prefrontal cortex. Journal of Neuroscience 24: 5704–10.CrossRefGoogle ScholarPubMed
Saper, C. B. (1985) Organization of cerebral cortical afferent systems in the rat. II. Hypothalamocortical projections. Journal of Comparative Neurology 237: 21–46.CrossRefGoogle ScholarPubMed
Saper, C. B. (1987) Diffuse cortical projection systems: anatomical organization and role in cortical function. In Handbook of Physiology, The Nervous System, sect. 1, vol. 5, ed. Moun, V. B. castle and Plum, F., pp. 169–210. Bethesda, MD: American Physiological Society.Google Scholar
Saper, C. B., Standaert, D. G., Currie, M. G., Schwartz, D., Geller, D. M. and Needleman, P. (1985) Atriopeptin-immunoreactive neurons in the brain: presence in cardiovascular regulatory areas. Science 227: 1047–9.CrossRefGoogle ScholarPubMed
Sarter, M. and Bruno, J. P. (2000) Cortical cholinergic inputs mediating arousal, attentional processing and dreaming: differential afferent regulation of the basal forebrain by telencephalic and brainstem afferents. Neuroscience 95: 933–52.CrossRefGoogle ScholarPubMed
Sarter, M. and Markowitsch, H. J. (1983) Convergence of basolateral amygdaloid and mediodorsal thalamic projections in different areas of the frontal cortex in the rat. Brain Research Bulletin 10: 607–22.CrossRefGoogle ScholarPubMed
Sarter, M. and Markowitsch, H. J. (1984) Collateral innervation of the medial and lateral prefrontal cortex by amygdaloid, thalamic, and brain-stem neurons. Journal of Comparative Neurology 224: 445–60.CrossRefGoogle ScholarPubMed
Satoh, K. and Fibiger, H. C. (1986) Cholinergic neurons of the laterodorsal tegmental nucleus: efferent and afferent connections. Journal of Comparative Neurology 253: 277–302.CrossRefGoogle ScholarPubMed
Saunders, M. G. and Westmoreland, B. F. (1979) The EEG in evaluation of disorders affecting the brain diffusely. In Current Practice of Clinical Electroencephalography, ed. Klass, D. W. and Daly, D. D., pp. 343–79. New York: Raven Press.Google Scholar
Saunders, R. C. and Rosene, D. L. (1988) A comparison of the efferents of the amygdala and the hippocampal formation in the rhesus monkey: I. Convergence in the entorhinal, prorhinal and perirhinal cortices. Journal of Comparative Neurology 271: 153–84.CrossRefGoogle ScholarPubMed
Saunders, R. C., Rosene, D. L. and Hoesen, G. W. (1988) Comparison of the efferents of the amygdala and the hippocampal formation in the rhesus monkey: II. Reciprocal and non-reciprocal connections. Journal of Comparative Neurology 271: 185–207.CrossRefGoogle ScholarPubMed
Savander, V., Ledoux, J. E. and Pitkänen, A. (1997) Interamygdaloid projections of the basal and accessory basal nuclei of the rat amygdaloid complex. Neuroscience 76: 725–35.CrossRefGoogle ScholarPubMed
Scalia, F. and Winans, S. S. (1975) The differential projections of the olfactory bulb and accessory olfactory bulb in mammals. Journal of Comparative Neurology 161: 31–56.CrossRefGoogle ScholarPubMed
Schafe, G. E., Nadel, N. V., Sullivan, G. M., Harris, A. and LeDoux, J. E. (1999) Memory consolidation for contextual and auditory fear conditioning is dependent on protein synthesis, PKA and MAP kinase. Learning and Memory 6: 97–110.Google ScholarPubMed
Schafe, G. E., Nader, K., Blair, H. T. and LeDoux, J. E. (2001) Memory consolidation of Pavlovian fear conditioning: a cellular and molecular perspective. Trends in Neurosciences 24: 540–6.CrossRefGoogle ScholarPubMed
Scharfman, H. E. (1991) Dentate hilar cells with dendrites in the molecular layer have lower thresholds for synaptic activation by perforant path than granule cells. Journal of Neuroscience 11: 1660–73.CrossRefGoogle ScholarPubMed
Schiess, M. C., Asprodini, E. K., Rainnie, D. G. and Shinnick-Gallagher, P. (1993) The central nucleus of the rat amygdala: in vitro intracellullar recordings. Brain Research 604: 283–97.CrossRefGoogle Scholar
Schiess, M. C., Callahan, P. M. and Zheng, H. (1999) Characterization of the electrophysiological and morphological properties of rat central amygdala neurons in vitro. Journal of Neuroscience Research 58: 663–73.3.0.CO;2-A>CrossRefGoogle ScholarPubMed
Schiff, N. D., Ribary, U., Moreno, D. R.et al. (2002) Residual cerebral activity and behavioral fragments can remain in the persistently vegetative brain. Brain 125: 1210–34.CrossRefGoogle ScholarPubMed
Schultz, W. (2002) Getting formal with dopamine and reward. Neuron 36: 241–63.CrossRefGoogle ScholarPubMed
Schwaber, J. S., Kapp, B. S., Higgins, G. A. and Rapp, P. R. (1982) Amygdaloid and basal forebrain direct connections with the nucleus of the solitary tract and the dorsal motor nucleus. Journal of Neuroscience 2: 1424–38.CrossRefGoogle ScholarPubMed
Schwindt, P. and Crill, W. (1999) Mechanisms underlying burst and regular spiking evoked by dendritic depolarization in layer 5 cortical pyramidal neurons. Journal of Neurophysiology 81: 1341–54.CrossRefGoogle ScholarPubMed
Schwindt, P. C., Spain, W. J., Foehring, R. C.et al. (1988a) Multiple potassium conductances and their functions in neurons from cat sensorimotor cortex in vitro. Journal of Neurophysiology 59: 424–49.CrossRefGoogle Scholar
Schwindt, P. C., Spain, W. J., Foehring, R. C., Chubb, M. C. and Crill, W. E. (1988b) Slow conductances in neurons from cat sensorimotor cortex in vitro and their role in slow excitability changes. Journal of Neurophysiology 59: 450–67.CrossRefGoogle Scholar
Schwindt, P. C., Spain, W. J. and Crill, W. E. (1989) Long-lasting reduction of excitability by a sodium-dependent potassium current in cat neocortical neurons. Journal of Neurophysiology 61: 233–44.CrossRefGoogle ScholarPubMed
Seamans, J. K., Nogueira, L. and Lavin, A. (2003) Synaptic basis of persistent activity in prefrontal cortex in vivo and in organotypic cultures. Cerebral Cortex 13: 1242–50.CrossRefGoogle ScholarPubMed
Segarra, J. M. (1970) Cerebral vascular disease and behaviour. I. The syndrome of the mesencephalic artery (basilary artery bifurcation). Archives of Neurology 22: 408–18.CrossRefGoogle Scholar
Seidemann, E., Meilijson, I., Abeles, M., Bergman, H. and Vaadia, E. (1996) Simultaneously recorded single units in the frontal cortex go through sequences of discrete and stable states in monkeys performing a delayed localization task. Journal of Neuroscience 16: 752–68.CrossRefGoogle ScholarPubMed
Seidenbecher, T., Laxmi, T. R., Stork, O. and Pape, H. C. (2003) Amygdalar and hippocampal theta rhythm synchronization during fear memory retrieval. Science 301: 846–50.CrossRefGoogle ScholarPubMed
Sejnowski, T. J. and Destexhe, A. (2000) Why do we sleep?Brain Research 886: 208–23.CrossRefGoogle ScholarPubMed
Semba, K. (1992) Development of central cholinergic neurons. In Handbook of Chemical Neuroanatomy, vol. 10 (Ontogeny and Transmitters of Peptides in the CNS), ed. Björkland, A., Hökfelt, T. and Tohyama, M., pp. 33–62. Amsterdam: Elsevier.Google Scholar
Semba, K. (2000) Multiple output pathways of the basal forebrain: organization, chemical heterogeneity, and role in vigilance. Behavioural Brain Research 115: 117–41.CrossRefGoogle ScholarPubMed
Semba, K. (2004) Phylogenetic and ontogenetic aspects of the basal forebrain cholinergic neurons and their innervation of the cerebral cortex. In Acetylcholine in the Cerebral Cortex (Progress in Brain Research, vol. 145), ed. Descarries, L., Krnjevic, K. and Steriade, M., pp. 3–43. Amsterdam: Elsevier.Google Scholar
Seroogy, K. B., Dangaran, K., Lim, S., Haycock, J. W. and Fallon, J. H. (1989) Ventral mesencephalic neurons containing both cholecystokinin- and tyrosine hydroxylase-like immunoreactivities project to forebrain regions. Journal of Comparative Neurology 279: 397–414.CrossRefGoogle ScholarPubMed
Sesack, S. R., Deutch, A. Y., Roth, R. H. and Bunney, B. S. (1989) Topographical organization of the efferent projections of the medial prefrontal cortex in the rat: an anterograde tract-tracing study with Phaseolus vulgaris leucoagglutinin. Journal of Comparative Neurology 290: 213–42.CrossRefGoogle ScholarPubMed
Shadlen, M. N. and Movshon, J. A. (1999) Synchrony unbound: a critical evaluation of the temporal binding hypothesis. Neuron 24: 67–77.CrossRefGoogle ScholarPubMed
Shaffery, J. P., Sinton, C. M., Bissette, G., Roffwarg, H. P. and Marks, G. A. (2002) Rapid eye movement sleep deprivation modifies expression of long-term potentiation in visual cortex of immature rats. Neuroscience 110: 431–43.CrossRefGoogle ScholarPubMed
Sheer, D. (1984) Focused arousal, 40 Hz, and dysfunction. In Selfregulation of the Brain and Behavior, ed. Ebert, T., pp. 64–84. Berlin: Springer.CrossRefGoogle Scholar
Shepherd, G. M. (2005) Perception without a thalamus: how does olfaction do it?Neuron 46: 166–8.CrossRefGoogle ScholarPubMed
Sherin, J. E., Shiromani, P. J., McCarley, R. W. and Saper, C. B. (1996) Activation of ventrolateral preoptic neurons during sleep. Science 271: 216–9.CrossRefGoogle ScholarPubMed
Sherin, J. E., Elmquist, J. K., Torrealba, F. and Saper, C. B. (1998) Innervation of histaminergic tuberomammillary neurons by GABAergic and galaninergic neurons in the ventrolateral preoptic nucleus of the rat. Journal of Neuroscience 18: 4705–21.CrossRefGoogle ScholarPubMed
Sherman, S. M. (2001) A wake-up call from the thalamus. Nature Neuroscience 4: 344–6.CrossRefGoogle ScholarPubMed
Sherrington, C. S. (1955) Man on his Nature. New York: Doubleday.Google Scholar
Shi, C. J. and Cassell, M. D. (1999) Perirhinal cortex projections to the amygdaloid complex and hippocampal formation in the rat. Journal of Comparative Neurology 406: 299–328.3.0.CO;2-9>CrossRefGoogle ScholarPubMed
Shi, S. H., Hayashi, Y., Petralia, R. S.et al. (1999) Rapid spine delivery and redistribution of AMPA receptors after synaptic NMDA receptor activation. Science 284: 1811–16.CrossRefGoogle ScholarPubMed
Shin, L. M., Orr, S. P., Carson, M. A.et al. (2004) Regional cerebral blood flow in the amygdala and medial prefrontal cortex during traumatic imagery in male and female Vietnam veterans with PTSD. Archives of General Psychiatry 61: 168–76.CrossRefGoogle ScholarPubMed
Shinonaga, Y., Takada, M. and Mizuno, N. (1994) Topographic organization of collateral projections from the basolateral amygdaloid nucleus to both the prefrontal cortex and nucleus accumbens in the rat. Neuroscience 58: 389–97.CrossRefGoogle ScholarPubMed
Shu, Y., Hasenstaub, A., Badoual, M., Bal, T. and McCormick, D. A. (2003) Barrages of synaptic activity control the gain and sensitivity of cortical neurons. Journal of Neuroscience 23: 10388–401.CrossRefGoogle ScholarPubMed
Siapas, A. G. and Wilson, M. A. (1998) Coordinated interactions between hippocampal ripples and cortical spindles during slow-wave sleep. Neuron 21: 1123–28.CrossRefGoogle ScholarPubMed
Siegel, M. and König, P. (2003) A functional gamma-band defined by stimulus-dependent synchronization in area 18 of awake behaving cats. Journal of Neuroscience 23: 4251–60.CrossRefGoogle ScholarPubMed
Sillito, A., Jones, H. E., Gerstein, G. L. and West, D. C. (1994) Feature-linked synchronization of thalamic relay cell firing induced by feedback from the visual cortex. Nature 369: 479–82.CrossRefGoogle ScholarPubMed
Silva, L. R., Gutnick, M. J. and Connors, B. W. (1991) Laminar distribution of neuronal membrane properties in neocortex of normal and reeler mouse. Journal of Neurophysiology 66: 2034–40.CrossRefGoogle ScholarPubMed
Simon, N. R., Mandshanden, I. and Lopes da Silva, F. H. (2000) A MEG study of sleep. Brain Research 860: 64–76.CrossRefGoogle Scholar
Singer, W. (1973) The effects of mesencephalic reticular stimulation on intracellular potentials of cat lateral geniculate neurons. Brain Research 61: 35–54.CrossRefGoogle Scholar
Singer, W. (1977) Control of thalamic transmission by corticofugal and ascending reticular pathways in the visual system. Physiological Reviews 57: 386–420.CrossRefGoogle ScholarPubMed
Singer, W. (1990) Role of acetylcholine in use-dependent plasticity of the visual cortex. In Brain Cholinergic Systems, ed. Steriade, M. and Biesold, D., pp. 314–36. Oxford: Oxford University Press.Google Scholar
Singer, W. (1999) Neuronal synchrony: a versatile code for the definition of relations?Neuron 24: 49–65.CrossRefGoogle ScholarPubMed
Sirota, A., Csicsvari, J., Buhl, D. and Buzsáki, G. (2003) Communication between neocortex and hippocampus during sleep in rodents. Proceedings of the National Academy of Sciences USA 100: 2065–9.CrossRefGoogle ScholarPubMed
Slotnick, S. D., Moo, L. R., Kraut, M. A., Lesser, R. P. and Hart, J. Jr. (2002) Interactions between thalamic and cortical rhythms during semantic memory recall in human. Proceedings of the National Academy of Sciences USA 99: 6440–3.CrossRefGoogle ScholarPubMed
Smith, Y. and Paré, D. (1994) Intra-amygdaloid projections of the lateral nucleus in the cat: PHA-L anterograde labeling combined with post-embedding GABA and glutamate immunocytochemistry. Journal of Comparative Neurology 342: 232–48.CrossRefGoogle Scholar
Smith, Y., Paré, D., Deschênes, M., Parent, A., and Steriade, M. (1988) Cholinergic and non-cholinergic projections from the upper brainstem core to the visual thalamus in the cat. Experimental Brain Research 70: 166–80.Google ScholarPubMed
Smith, Y., Paré, J. F. and Paré, D. (1998) Cat intraamygdaloid inhibitory network: ultrastructural organization of parvalbumin-immunoreactive elements. Journal of Comparative Neurology 391: 164–79.3.0.CO;2-0>CrossRefGoogle ScholarPubMed
Smith, Y., Paré, J. F. and Paré, D. (2000) Differential innervation of parvalbumin-immunoreactive interneurons of the basolateral amygdaloid complex by cortical and intrinsic inputs. Journal of Comparative Neurology 416: 496–508.3.0.CO;2-N>CrossRefGoogle ScholarPubMed
Snowden, R. J., Treue, S. and Andersen, R. A. (1992) The response of neurons in areas V1 and MT of the alert rhesus monkey to moving random dot patterns. Experimental Brain Research 88: 389–400.CrossRefGoogle ScholarPubMed
Sobotka, S. and Ringo, J. L. (1993) Investigation of long-term recognition and association memory in unit responses from inferotemporal cortex. Experimental Brain Research 96: 28–38.CrossRefGoogle ScholarPubMed
Soderling, T. R. and Derkach, V. A. (2000) Postsynaptic protein phosphorylation and LTP. Trends in Neurosciences 23: 75–80.CrossRefGoogle ScholarPubMed
Soltesz, I. and Crunelli, V. (1992) GABAA and pre- and post-synaptic GABAB receptor-mediated responses in the lateral geniculate nucleus. In Progress in Brain Research, vol. 90, ed. Mize, R. R., Marc, R. E. and Sillito, A. M., pp. 151–69. Amsterdam: Elsevier.Google Scholar
Soltesz, I. and Deschênes, M. (1993) Low- and high-frequency membrane potential oscillations during theta activity in CA1 and CA3 pyramidal neurons of the rat hippocampus under ketamine-xylazine anesthesia. Journal of Neurophysiology 70: 97–116.CrossRefGoogle ScholarPubMed
Soltesz, I., Lightowler, S., Leresche, N., Jassik-Gerschenfeld, D. and Crunelli, V. (1991) Two inward currents and the transformation of low-frequency oscillations of rat and cat thalamocortical cells. Journal of Physiology (London) 441: 175–97.CrossRefGoogle ScholarPubMed
Somogyi, P. (1977) A specific ‘axo-axonal’ interneuron in the visual cortex of the rat. Brain Research 136: 345–50.CrossRefGoogle ScholarPubMed
Somogyi, P. (1978) The study of Golgi stained cells and the experimental degeneration under the electron microscope: a direct method for the identification in the visual cortex of three successive links in a neuronal chain. Neuroscience 3: 167–80.CrossRefGoogle Scholar
Somogyi, P. and Cowey, A. (1981) Combined Golgi and electron microscopic study on the synapses formed by double bouquet cells in the visual cortex of the cat and monkey. Journal of Comparative Neurology 195: 547–66.CrossRefGoogle ScholarPubMed
Somogyi, P., Kisvárday, Z. F., Martin, K. A. C. and Whitteridge, D. (1983) Synaptic connections of morphologically identified and physiologically characterized large basket cells in the striate cortex of cat. Neuroscience 10: 261–94.CrossRefGoogle ScholarPubMed
Somogyi, P., Freund, T. F., Hodgson, A. J.et al. (1985) Identified axo-axonic cells are immunoreactive for GABA in the hippocampus and visual cortex of cats. Brain Research 332: 143–9.CrossRefGoogle Scholar
Somogyi, P., Tamás, G., Lujan, R. and Buhl, E. H. (1998) Salient features of synaptic organisation in the cerebral cortex. Brain Research Reviews 26: 113–35.CrossRefGoogle ScholarPubMed
Soriano, E., Martinez, A., Fariñas, I. and Frotscher, M. (1993) Chandelier cells in the hippocampal formation of the rat: the entorhinal area and subicular complex. Journal of Comparative Neurology 337: 151–67.CrossRefGoogle ScholarPubMed
Sorvari, H., Soininen, H., Paljärvi, L., Karkola, K. and Pitkänen, A. (1995).Distribution of parvalbumin-immunoreactive cells and fibers in the human amygdaloid complex. Journal of Comparative Neurology 360: 185–212.CrossRefGoogle ScholarPubMed
Soury, J. (1899) Le Système Nerveux Central. Histoire Critique des Théories et des Doctrines. Paris: Carré et Naud.Google Scholar
Spiegel, E. (1919) Die Kerne im Voderhirn des Sauger. Arbeiten der Neurologische Institut – Wien Universität 22: 418–97.Google Scholar
Spreafico, R., Amadeo, A., Angoscini, P., Panzica, F. and Battaglia, G. (1993) Branching projections from mesopontine nuclei to the nucleus reticularis and related thalamic nuclei: a double labelling study in the rat. Journal of Comparative Neurology 336: 481–92.CrossRefGoogle ScholarPubMed
Squire, L. R. and Alvarez, P. (1995) Retrograde amnesia and memory consolidation: a neurobiological perspective. Current Opinions in Neurobiology 5: 169–77.CrossRefGoogle ScholarPubMed
Standaert, D. G., Saper, C. B., Rye, D. B. and Wainer, B. H. (1986) Colocalization of atriopeptin-like immunoreactivity with choline acetyltransferase and substance P-like immunoreactivity in the pedunculopontine and laterodorsal tegmental nuclei in the rat. Brain Research 382: 163–8.CrossRefGoogle ScholarPubMed
Stanford, L. R., Friedländer, M. J. and Sherman, S. M. (1983) Morphological and physiological properties of geniculate W-cells of the cat: a comparison with X- and Y-cells. Journal of Neurophysiology 50: 582–608.CrossRefGoogle ScholarPubMed
Stefanacci, L., Suzuki, W. A. and Amaral, D. G. (1996) Organization of connections between the amygdaloid complex and the perirhinal and parahippocampal cortices in macaque monkeys. Journal of Comparative Neurology 375: 552–82.3.0.CO;2-0>CrossRefGoogle ScholarPubMed
Steininger, T. L., Gong, H., McGinty, D. and Szymusiak, R. (2001) Subregional organization of preoptic area/anterior hypothalamic projections to arousal-related monoaminergic cell groups. Journal of Comparative Neurology 429: 638–53.3.0.CO;2-Y>CrossRefGoogle ScholarPubMed
Steriade, M. (1970) Ascending control of thalamic and cortical responsiveness. International Review of Neurobiology 12: 87–144.CrossRefGoogle ScholarPubMed
Steriade, M. (1976) Cortical inhibition during sleep and waking. In Mechanisms in Transmission of Signal for Conscious Behavior, ed. Desiraju, T., pp. 209–48. Amsterdam: Elsevier.Google Scholar
Steriade, M. (1978) Cortical long-axoned cells and putative interneurons during the sleep-waking cycle. Behavioral and Brain Sciences 3: 465–514.CrossRefGoogle Scholar
Steriade, M. (1981) Mechanisms underlying cortical activation: neuronal organization and properties of the midbrain reticular core and intralaminar thalamic nuclei. In Brain Mechanisms and Perceptual Awareness, ed. Pompeiano, O. and Ajmone-Marsan, C., pp. 327–77. New York: Raven.Google Scholar
Steriade, M. (1984) The excitatory-inhibitory response sequence of thalamic and neocortical cells: state-related changes and regulatory systems. In Dynamic Aspects of Neocortical Function, ed. Edelman, G. M., Gall, W. E. and Cowan, W. M., pp. 107–57. New York: Wiley.Google Scholar
Steriade, M. (1991) Alertness, quiet sleep, dreaming. In Cerebral Cortex, vol. 9 (Normal and Altered States of Function), ed. Peters, A. and Jones, E. G., pp. 279–357. New York: Plenum.CrossRefGoogle Scholar
Steriade, M. (1995a) Two channels in the cerebellothalamocortical system. Journal of Comparative Neurology 354: 57–70.CrossRefGoogle Scholar
Steriade, M. (1995b) Brain activation, then (1949) and now: coherent fast rhythms in corticothalamic networks. Archives Italiennes de Biologie 134: 5–20.Google Scholar
Steriade, M. (1997a) Synchronized activities of coupled oscillators in the cerebral cortex and thalamus at different levels of vigilance. Cerebral Cortex 7: 583–604.CrossRefGoogle Scholar
Steriade, M. (1997b) Thalamic substrates of disturbances in states of vigilance and consciousness in humans. In Thalamus, vol. 2 (Experimental and Clinical Aspects), ed. Steriade, M., Jones, E. G. and McCormick, D. A., pp. 721–42. Oxford: Elsevier.Google Scholar
Steriade, M. (1999) Coherent oscillations and short-term plasticity in corticothalamic networks. Trends in Neurosciences 22: 337–45.CrossRefGoogle ScholarPubMed
Steriade, M. (2000) Corticothalamic resonance, states of vigilance, and mentation. Neuroscience 101: 243–76.CrossRefGoogle ScholarPubMed
Steriade, M. (2001a) The Intact and Sliced Brain. Cambridge, MA: MIT Press.Google Scholar
Steriade, M. (2001b) Impact of network activities on neuronal properties in corticothalamic systems. Journal of Neurophysiology 86: 1–39.CrossRefGoogle Scholar
Steriade, M. (2001c) The GABAergic reticular nucleus: a preferential target of corticothalamic projections. Proceedings of the National Academy of Sciences USA 98: 3625–7.CrossRefGoogle Scholar
Steriade, M. (2001d) To burst, or rather, not to burst. Nature Neuroscience 4: 671.CrossRefGoogle Scholar
Steriade, M. (2003a) Neuronal Substrates of Sleep and Epilepsy. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Steriade, M. (2003b) The corticothalamic system in sleep. Frontiers in Bioscience 8: 878–99.CrossRefGoogle Scholar
Steriade, M. (2003c) Presynaptic dendrites of thalamic local-circuit neurons and sculpting inhibition during activated states. Journal of Physiology (London) 546: 1.CrossRefGoogle Scholar
Steriade, M. (2004a) Neocortical neuronal classes are flexible entities. Nature Reviews Neuroscience 5: 121–34.CrossRefGoogle Scholar
Steriade, M. (2004b).Local gating of information processing through the thalamus. Neuron 41: 493–4.CrossRefGoogle Scholar
Steriade, M. (2004c).Slow-wave sleep: serotonin, neuronal plasticity, and seizures. Archives Italiennes de Biologie 142: 359–67.Google Scholar
Steriade, M. (2005a) Cellular substrates of brain rhythms. In Electroencephalography: Basic Principles, Clinical Applications, and Related Fields (5th edition), ed. Niedermeyer, E. and Silva, F. Lopes Da, pp. 31–83. Baltimore, MD: Williams & Wilkins.Google Scholar
Steriade, M. (2005b) Neuronal substrates of spike-wave seizures and hypsarrhythmia in corticothalamic systems. Advances in Neurology 27: 149–54.Google Scholar
Steriade, M. (2006) Brainstem-thalamic neurons implicated in hallucinations. Behavioral and Brain Sciences, in press.Google Scholar
Steriade, M. and Amzic, F. (1996) Intracortical and corticothalamic coherency of fast spontaneous oscillations. Proceedings of the National Academy of Sciences USA 93: 2533–8.CrossRefGoogle ScholarPubMed
Steriade, M. and Amzica, F. (1998) Coalescence of sleep rhythms and their chronology in corticothalamic networks. Sleep Research Online 1: 1–10.Google ScholarPubMed
Steriade, M. and Buzsáki, G. (1990). Parallel activation of thalamic and cortical neurons by brainstem and basal forebrain cholinergic systems. In Brain Cholinergic Systems, ed. Steriade, M. and Biesold, D., pp. 3–63. Oxford: Oxford University Press.Google Scholar
Steriade, M. and Contreras, D. (1993) Sleep oscillations in interacting thalamocortical networks. In Thalamic Networks for Relay and Modulation, ed. Minciacchi, D., Molinari, M., Macchi, G. and Jones, E. G., pp. 385–94. Oxford: Pergamon Press.Google Scholar
Steriade, M. and Contreras, D. (1995) Relations between cortical and thalamic cellular events during transition from sleep pattern to paroxysmal activity. Journal of Neuroscience 15: 623–42.CrossRefGoogle Scholar
Steriade, M. and Demetrescu, M. (1960) Unspecific systems of inhibition and facilitation of potentials evoked by intermittent light. Journal of Neurophysiology 23: 602–17.CrossRefGoogle Scholar
Steriade, M. and Demetrescu, M. (1967) Simulation of peripheral sensory input by electrical pulse trains applied to specific afferent pathways. Experimental Neurology 19: 265–77.CrossRefGoogle ScholarPubMed
Steriade, M. and Deschênes, M. (1974) Inhibitory processes and interneuronal apparatus in motor cortex during sleep and waking. II. Recurrent and afferent inhibition of pyramidal tract neurons. Journal of Neurophysiology 37: 1093–113.CrossRefGoogle ScholarPubMed
Steriade, M. and Deschênes, M. (1984) The thalamus as a neuronal oscillator. Brain Research Reviews 8: 1–63.CrossRefGoogle Scholar
Steriade, M. and Deschênes, M. (1987) Inhibitory processes in the thalamus. Journal of Mind and Behavior 8: 559–72.Google Scholar
Steriade, M. and Deschênes, M. (1988) Intrathalamic and brainstem-thalamic networks involved in resting and alert states. In Cellular Thalamic Mechanisms, ed. Bentivoglio, M., Macchi, G. and Spreafico, R., pp. 37–62. Amsterdam: Elsevier.Google Scholar
Steriade, M. and Glenn, L. L. (1982) Neocortical and caudate projections of intralaminar thalamic neurons and their synaptic excitation from the midbrain reticular core. Journal of Neurophysiology 48: 352–71.CrossRefGoogle ScholarPubMed
Steriade, M. and Hobson, J. A. (1976) Neuronal activity during the sleep-waking cycle. Progress in Neurobiology 6: 155–376.CrossRefGoogle ScholarPubMed
Steriade, M. and Llinás, R. R. (1988) The functional states of the thalamus and the associated neuronal interplay. Physiological Reviews 68: 649–742.CrossRefGoogle ScholarPubMed
Steriade, M. and McCarley, R. W. (2005) Brain Control of Wakefulness and Sleep (2nd edition). New York: Springer-Kluwer.Google Scholar
Steriade, M. and Morin, D. (1981) Reticular influences on primary and augmenting responses in the somatosensory cortex. Brain Research 205: 67–80.CrossRefGoogle ScholarPubMed
Steriade, M. and Timofeev, I. (1997) Short-term plasticity during intrathalamic augmenting responses in decorticated cats. Journal of Neuroscience 17: 3778–95.CrossRefGoogle ScholarPubMed
Steriade, M. and Timofeev, I. (2001) Corticothalamic operations through prevalent inhibition of thalamocortical neurons. Thalamus and Related Systems 1: 225–36.Google Scholar
Steriade, M. and Timofeev, I. (2003) Neuronal plasticity in thalamocortical networks during sleep and waking oscillations. Neuron 37: 563–76.CrossRefGoogle ScholarPubMed
Steriade, M., Belekhova, M. and Apostol, V. (1968) Reticular potentiation of cortical flash-evoked afterdischarge. Brain Research 11: 276–80.CrossRefGoogle ScholarPubMed
Steriade, M., Iosif, G. and Apostol, V. (1969) Responsiveness of thalamic and cortical motor relays during arousal and various stages of sleep. Journal of Neurophysiology 32: 251–65.CrossRefGoogle Scholar
Steriade, M., Apostol, V. and Oakson, G. (1971) Control of unitary activities in cerebellothalamic pathway during wakefulness and synchronized sleep. Journal of Neurophysiology 34: 389–413.CrossRefGoogle ScholarPubMed
Steriade, M., Wyzinski, P. and Apostol, V. (1972) Corticofugal projections governing rhythmic thalamic activity. In Corticothalamic Projections and Sensorimotor Activities, ed. Frigyesi, T. L., Rinvik, E. and Yahr, M. D., pp. 221–72. New York: Raven Press.Google Scholar
Steriade, M., Deschênes, M. and Oakson, G. (1974a) Inhibitory processes and interneuronal apparatus in motor cortex during sleep and waking. I. Background firing and synaptic responsiveness of pyramidal tract neurons and interneurons. Journal of Neurophysiology 37: 1065–92.CrossRefGoogle Scholar
Steriade, M., Deschênes, M., Wyzinski, P. and Hallé, J. P. (1974b) Input-output organization of the motor cortex during sleep and waking. In Basic Sleep Mechanisms, ed. Petre-Quadens, O. and Schlag, J., pp. 144–200. New York: Academic Press.Google Scholar
Steriade, M., Oakson, G. and Diallo, A. (1977) Reticular influences on lateralis posterior thalamic neurons. Brain Research 131: 55–71.CrossRefGoogle ScholarPubMed
Steriade, M., Kitsikis, A. and Oakson, G. (1979) Excitatory-inhibitory processes in parietal association neurons during reticular activation and sleep-waking cycle. Sleep 1: 339–55.CrossRefGoogle ScholarPubMed
Steriade, M., Oakson, G. and Ropert, N. (1982a) Firing rates and patterns of midbrain reticular neurons during steady and transitional states of the sleep-waking cycle. Experimental Brain Research 46: 37–51.CrossRefGoogle Scholar
Steriade, M., Parent, A., Ropert, N. and Kitsikis, A. (1982b) Zona incerta and lateral hypothalamic afferents to the midbrain reticular core of the cat – an HRP and electrophysiological study. Brain Research 238: 13–28.CrossRefGoogle Scholar
Steriade, M., Parent, A. and Hada, J. (1984a) Thalamic projections of nucleus reticularis thalami of cat: a study using retrograde transport of horseradish peroxidase and double fluorescent tracers. Journal of Comparative Neurology 229: 531–47.CrossRefGoogle Scholar
Steriade, M., Sakai, K. and Jouvet, M. (1984b) Bulbothalamic neurons related to thalamocortical activation processes during paradoxical sleep. Experimental Brain Research 54: 463–75.CrossRefGoogle Scholar
Steriade, M., Deschênes, M., Domich, L. and Mulle, C. (1985) Abolition of spindle oscillations in thalamic neurons disconnected from nucleus reticularis thalami. Journal of Neurophysiology 54: 1473–97.CrossRefGoogle ScholarPubMed
Steriade, M., Domich, L. and Oakson, G. (1986) Reticularis thalami neurons revisited: activity changes during shifts in states of vigilance. Journal of Neuroscience 6: 68–81.CrossRefGoogle ScholarPubMed
Steriade, M., Domich, L., Oakson, G. and Deschênes, M. (1987a) The deafferented reticular thalamic nucleus generates spindle rhythmicity. Journal of Neurophysiology 57: 260–73.CrossRefGoogle Scholar
Steriade, M., Parent, A., Paré, D. and Smith, Y. (1987b) Cholinergic and non-cholinergic neurons of cat basal forebrain project to reticular and mediodorsal thalamic nuclei. Brain Research 408: 372–76.CrossRefGoogle Scholar
Steriade, M., Paré, D., Parent, A. and Smith, Y. (1988) Projections of cholinergic and non-cholinergic neurons of the brainstem core to relay and associational thalamic nuclei in the cat and macaque monkey. Neuroscience 25: 47–67.CrossRefGoogle ScholarPubMed
Steriade, M., Paré, D., Bouhassira, D., Deschênes, M. and Oakson, G. (1989) Phasic activation of lateral geniculate and perigeniculate neurons during sleep with ponto-geniculo-occipital spikes. Journal of Neuroscience 9: 2215–29.CrossRefGoogle Scholar
Steriade, M., Datta, S., Paré, D., Oakson, G. and Curró Dossi, R. (1990a) Neuronal activities in brainstem cholinergic nuclei related to tonic activation processes in thalamocortical systems. Journal of Neuroscience 10: 2541–59.CrossRefGoogle Scholar
Steriade, M., Jones, E. G. and Llinás, R. R. (1990b) Thalamic Oscillations and Signaling. New York: Wiley-Interscience.Google Scholar
Steriade, M., Gloor, P., Llinás, R. R., Lopes da Silva, F. H. and Mesulam, M. M. (1990c) Basic mechanisms of cerebral rhythmic activities. Electroencephalography and Clinical Neurophysiology 76: 481–508.CrossRefGoogle Scholar
Steriade, M., Paré, D., Datta, S., Oakson, G. and Curró Dossi, R. (1990d) Different cellular types in mesopontine cholinergic nuclei related to ponto-geniculo-occipital waves. Journal of Neuroscience 10: 2560–79.CrossRefGoogle Scholar
Steriade, M., Curró Dossi, R. and Nuñez, A. (1991a) Network modulation of a slow intrinsic oscillation of cat thalamocortical neurons implicated in sleep delta waves: cortical potentiation and brainstem cholinergic suppression. Journal of Neuroscience 11: 3200–17.CrossRefGoogle Scholar
Steriade, M., Curró Dossi, R., Paré, D. and Oakson, G. (1991b) Fast oscillations (20–40 Hz) in thalamocortical systems and their potentiation by mesopontine cholinergic nuclei in the cat. Proceedings of the National Academy of Sciences USA 88: 4396–400.CrossRefGoogle Scholar
Steriade, M., Amzica, F. and Nuñez, A. (1993a) Cholinergic and noradrenergic modulation of the slow (∼0.3 Hz) oscillation in neocortical cells. Journal of Neurophysiology 70: 1384–400.CrossRefGoogle Scholar
Steriade, M., Contreras, D., Curró Dossi, R. and Nuñez, A. (1993b) The slow (<1 Hz) oscillation in reticular thalamic and thalamocortical neurons: scenario of sleep rhythm generation in interacting thalamic and neocortical networks. Journal of Neuroscience 13: 3284–99.CrossRefGoogle Scholar
Steriade, M., Curró Dossi, R. and Contreras, D. (1993c) Electrophysiological properties of intralaminar thalamocortical cells discharging rhythmic (∼40 Hz) spike-bursts at ∼1000 Hz during waking and rapid eye movement sleep. Neuroscience 56: 1–9.CrossRefGoogle Scholar
Steriade, M., McCormick, D. A. and Sejnowski, T. J. (1993d) Thalamocortical oscillation in the sleeping and aroused brain. Science 262: 679–85.CrossRefGoogle Scholar
Steriade, M., Nuñez, A. and Amzica, F. (1993e) A novel slow (<1 Hz) oscillation of neocortical neurons in vivo: depolarizing and hyperpolarizing components. Journal of Neuroscience 13: 3252–65.CrossRefGoogle Scholar
Steriade, M., Nuñez, A. and Amzica, F. (1993f) Intracellular analysis of relations between the slow (<1 Hz) neocortical oscillation and other sleep rhythms. Journal of Neuroscience 13: 3266–83.CrossRefGoogle Scholar
Steriade, M., Amzica, F. and Contreras, D. (1994a) Cortical and thalamic cellular correlates of electroencephalographic burst-suppression. Electroencephalography and Clinical Neurophysiology 90: 1–16.CrossRefGoogle Scholar
Steriade, M., Contreras, D. and Amzica, F. (1994b) Synchronized sleep oscillations and their paroxysmal developments. Trends in Neuroscience 17: 199–208.CrossRefGoogle Scholar
Steriade, M., Amzica, F. and Contreras, D. (1996a) Synchronization of fast (30–40 Hz) spontaneous cortical rhythms during brain activation. Journal of Neuroscience 16: 392–417.CrossRefGoogle Scholar
Steriade, M., Contreras, D., Amzica, F. and Timofeev, I. (1996b) Synchronization of fast (30–40 Hz) spontaneous oscillations in intrathalamic and thalamocortical networks. Journal of Neuroscience 16: 2788–808.CrossRefGoogle Scholar
Steriade, M., Jones, E. G. and McCormick, D. A. (1997) Thalamus, vol. 1 (Organisation and Function). Oxford: Elsevier.Google Scholar
Steriade, M., Timofeev, I., Dürmüller, N. and Grenier, F. (1998a) Dynamic properties of corticothalamic neurons and local cortical interneurons generating fast rhythmic (30–40 Hz) spike bursts. Journal of Neurophysiology 79: 483–90.CrossRefGoogle Scholar
Steriade, M., Timofeev, I., Grenier, F. and Dürmüller, N. (1998b) Role of thalamic and cortical neurons in augmenting responses: dual intracellular recordings in vivo. Journal of Neuroscience 18: 6425–43.CrossRefGoogle Scholar
Steriade, M., Timofeev, I. and Grenier, F. (2001a) Natural waking and sleep states: a view from inside neocortical neurons. Journal of Neurophysiology 85: 1969–85.CrossRefGoogle Scholar
Steriade, M., Timofeev, I. and Grenier, F. (2001b) Intrinsic, antidromic and synaptic excitability of cortical neurons during natural waking-sleep cycle. Society for Neuroscience Abstracts 27: 240.Google Scholar
Stickgold, R., James, L. and Hobson, J. A. (2000a) Visual discrimination learning requires sleep after training. Nature Neuroscience 3: 1237–8.CrossRefGoogle Scholar
Stickgold, R., Whitbee, D., Schirmer, B., Patel, V. and Hobson, J. A. (2000b) Visual discrimination improvement. A multi-step process occurring during sleep. Journal of Cognitive Neuroscience 12: 246–54.CrossRefGoogle Scholar
Stone, V. E., Baron-Cohen, S. and Knight, R. T. (1998) Frontal lobe contributions in theory of the mind. Journal of Cognitive Neuroscience 10: 640–56.CrossRefGoogle ScholarPubMed
Storm, J. F. (1988) Temporal integration by a slowly inactivating K+ current in hippocampal neurons. Nature 336: 379–81.CrossRefGoogle ScholarPubMed
Stratford, K. J., Tarczy-Hornoch, K., Martin, K. A. C., Bannister, N. J. and Jack, J. J. B. (1996) Excitatory synaptic inputs to spiny stellate cells in cat visual cortex. Nature 382: 258–60.CrossRefGoogle ScholarPubMed
Stuart, G. and Sakmann, B. (1994) Active propagation of somatic action potentials into neocortical pyramidal cell dendrites. Nature 367: 69–72.CrossRefGoogle ScholarPubMed
Stuart, G., Spruston, N., Sakmann, B. and Häusser, M. (1997) Action potential initiation and backpropagation in neurons of the mammalian CNS. Trends in Neurosciences 20: 125–31.CrossRefGoogle ScholarPubMed
Sugihara, I., Lang, E. J. and Llinás, R. (1993) Uniform olivocerebellar conduction time underlies Purkinje cell complex spike synchronicity in the rat cerebellum. Journal of Physiology (London) 470: 243–71.CrossRefGoogle ScholarPubMed
Sun, N. and Cassell, M. D. (1993) Intrinsic GABAergic neurons in the rat central extended amygdala. Journal of Comparative Neurology 330: 381–404.CrossRefGoogle ScholarPubMed
Sutton, R. E., Koob, G. F., Moal, M., Rivier, J. and Vale, W. (1982) Corticotropin releasing factor produces behavioral activation in rats. Nature 297: 31–3.CrossRefGoogle ScholarPubMed
Suzuki, W. A. (1996) The anatomy, physiology and functions of the perirhinal cortex. Current Opinion in Neurobiology 6: 179–86.CrossRefGoogle ScholarPubMed
Suzuki, W. A. and Porteros, A. (2002) Distribution of calbindin D-28k in the entorhinal, perirhinal, and parahippocampal cortices of the macaque monkey. Journal of Comparative Neurology 451: 392–412.CrossRefGoogle ScholarPubMed
Suzuki, W. A., Zola-Morgan, S., Squire, L. R. and Amaral, D. G. (1993) Lesions of the perirhinal and parahippocampal cortices in the monkey produce long-lasting memory impairment in the visual and tactual modalities. Journal of Neuroscience 13: 2430–51.CrossRefGoogle ScholarPubMed
Svoboda, K., Denk, W., Kleinfeld, D. and Tank, D. W. (1997) In vivo dendritic calcium dynamics in neocortical neurons. Nature 385: 161–5.CrossRefGoogle Scholar
Svoboda, K., Helmchen, F., Denk, W. and Tank, D. W. (1999) Spread of dendritic excitation in layer 2/3 pyramidal neurons in rat barrel cortex in vivo. Nature Neuroscience 2: 65–73.CrossRefGoogle ScholarPubMed
Swadlow, H. A. and Gusev, A. G. (2001) The impact of ‘bursting’ thalamic impulses at a neocortical synapse. Nature Neuroscience 4: 402–8.CrossRefGoogle Scholar
Swanson, L. W. (1992) Brain Maps: Structure of the Rat Brain. Amsterdam: Elsevier.Google Scholar
Swanson, L. W. and Hartman, B. K. (1975) The central adrenergic system. An immunofluorescence study of the location of cell bodies and their efferent connections in the rat utilizing dopamine-beta-hydroxylase as a marker. Journal of Comparative Neurology 163: 467–505.CrossRefGoogle ScholarPubMed
Swanson, L. W. and Petrovich, G. D. (1998) What is the amygdala?Trends in Neurosciences 21: 323–31.CrossRefGoogle ScholarPubMed
Swanson, L. W., Mogenson, G. J., Simerly, R. B. and Wu, M. (1987) Anatomical and electrophysiological evidence for a projection from the medial preoptic area to the ‘mesencephalic and subthalamic locomotor regions’ in the rat. Brain Research 405: 108–22.CrossRefGoogle ScholarPubMed
Szymusiak, R. and McGinty, D. (1986) Sleep-related neuronal discharge in the basal forebrain of cats. Brain Research 370: 82–92.CrossRefGoogle ScholarPubMed
Szymusiak, R., Steininger, T., Alam, N. and McGinty, D. (2001) Preoptic area sleep-regulating mechanisms. Archives Italiennes de Biologie 139: 77–92.Google ScholarPubMed
Takada, M. (1990) The A11 catecholamine cell group: another origin of the dopaminergic innervation of the amygdala. Neuroscience Letters 118: 132–5.CrossRefGoogle ScholarPubMed
Takagishi, M. and Chiba, T. (1991) Efferent projections of the infralimbic (area 25) region of the medial prefrontal cortex in the rat: an anterograde tracer PHA-L study. Brain Research 566: 26–36.CrossRefGoogle ScholarPubMed
Takakusaki, K. and Kitai, S. T. (1997) Ionic mechanisms involved in the spontaneous firing of tegmental pedunculopontine nucleus neurons of the rat. Neuroscience 78: 771–94.CrossRefGoogle ScholarPubMed
Takakusaki, K., Shiroyama, T. and Kitai, S. T. (1997) Two types of cholinergic neurons in he rat tegmental pedunculopontine nucleus: electrophysiological and morphological characterization. Neuroscience 79: 1089–99.CrossRefGoogle Scholar
Takeuchi, Y., McLean, J. H. and Hopkins, D. A. (1982) Reciprocal connections between the amygdala and parabrachial nuclei: ultrastructural demonstration by degeneration and axonal transport of horse radish peroxidase in the cat. Brain Research 239: 583–8.CrossRefGoogle Scholar
Takeuchi, Y., Matsushima, S., Matsuchima, R. and Hopkins, D. A. (1983) Direct amygdaloid projections to the dorsal motor nucleus of the vagus nerve: a light and electron microscopic study in the rat. Brain Research 280: 143–7.CrossRefGoogle ScholarPubMed
Takita, M., Izaki, Y., Jay, T. M., Kaneko, H. and Suzuki, S. S. (1999) Induction of stable long-term depression in vivo in the hippocampal-prefrontal cortex pathway. European Journal of Neuroscience 11: 4145–8.CrossRefGoogle ScholarPubMed
Tamás, G., Buhl, E. H. and Somogyi, P. (1997) Massive autaptic self-innervation of GABAergic neurons in cat visual cortex. Journal of Neuroscience 17: 6352–64.CrossRefGoogle ScholarPubMed
Tamás, G., Somogyi, P. and Buhl, E. H. (1998) Differentially interconnected networks of GABAergic interneurons in the visual cortex of the cat. Journal of Neuroscience 18: 4255–70.CrossRefGoogle ScholarPubMed
Tanabe, T., Yarita, H., Iino, M., Ooshima, Y. and Takagi, S. F. (1975a) An olfactory projection area in orbitofrontal cortex of the monkey. Journal of Neurophysiology 38: 1269–83.CrossRefGoogle Scholar
Tanabe, T., Iino, M. and Takagi, S. F. (1975b) Discrimination of odors in olfactory bulb, pyriform-amygdaloid areas, and orbitofrontal cortex of the monkey. Journal of Neurophysiology 38: 1284–96.CrossRefGoogle Scholar
Tancredi, V., Biagini, G., D'Antuono, M., Louvel, J., Pumain, R. and Avoli, M. (2000) Spindle-like thalamocortical synchronization in a rat brain slice preparation. Journal of Neurophysiology 84: 1093–7.CrossRefGoogle Scholar
Tang, A. C., Bartels, A. M. and Sejnowski, T. J. (1997) Effects of cholinergic modulation on responses of neocortical neurons to fluctuating input. Cerebral Cortex 7: 502–9.CrossRefGoogle ScholarPubMed
Terreberry, R. R. and Neafsey, E. J. (1987) The rat medial frontal cortex projects directly to autonomic regions of the brainstem. Brain Research Bulletin 19: 639–49.CrossRefGoogle ScholarPubMed
Terzano, M. G., Parrino, L. and Spaggiari, M. C. (1988) The cyclic alternating pattern sequences in the dynamic organization of sleep. Electroencephalography and Clinical Neurophysiology 69: 437–47.CrossRefGoogle Scholar
Thierry, A. M., Gioanni, Y., Degenetais, E. and Glowinski, J. (2000) Hippocampo-prefrontal cortex pathway: anatomical and electrophysiological characteristics. Hippocampus 10: 411–19.3.0.CO;2-A>CrossRefGoogle ScholarPubMed
Thomson, A. M. (1988a) Inhibitory postsynaptic potentials evoked in thalamic neurons by stimulation of the reticularis nucleus evoke slow spikes in isolated rat brain slices. Neuroscience 25: 491–502.CrossRefGoogle Scholar
Thomson, A. M. (1988b) Biphasic responses of thalamic neurons to GABA in isolated rat brain slices. Neuroscience 25: 503–12.CrossRefGoogle Scholar
Thomson, A. M. (1997) Activity-dependent properties of synaptic transmission at two classes of connections made by rat neocortical pyramidal neurons in vitro. Journal of Physiology (London) 502: 131–47.CrossRefGoogle ScholarPubMed
Thomson, A. M. and Deuchars, J. (1997) Synaptic interactions in neocortical local circuits: dual intracellular recordings in vitro. Cerebral Cortex 7: 510–22.CrossRefGoogle ScholarPubMed
Thomson, A. M. and Morris, O. T. (2002) Selectivity in the inter-laminar connections made by neocortical neurones. Journal of Neurocytology 31: 239–46.CrossRefGoogle ScholarPubMed
Thomson, A. M., West, D. C., Hahn, J. and Deuchars, J. (1996) Single axon IPSPs elicited in pyramidal cells by three classes of interneurons in slices of rat neocortex. Journal of Physiology (London) 496: 81–102.CrossRefGoogle ScholarPubMed
Thomson, A. M., West, D. C., Wang, Y. and Bannister, A. P. (2002) Synaptic connections and small circuits involving excitatory and inhibitory neurons in layers 2–5 of adult rat and cat neocortex: triple intracellular recordings and biocytin labelling in vitro. Cerebral Cortex 12: 936–53.CrossRefGoogle ScholarPubMed
Tigges, J., Tigges, M., Cross, N. A.et al. (1982) Subcortical structures projecting to visual cortical areas in squirrel monkey. Journal of Comparative Neurology 209: 29–40.CrossRefGoogle ScholarPubMed
Tigges, J., Walker, L. C. and Tigges, M. (1983) Subcortical projections to the occipital lobe and parietal lobes of the chimpanzee brain. Journal of Comparative Neurology 220: 106–15.CrossRefGoogle ScholarPubMed
Timofeev, I. and Steriade, M. (1996) Low-frequency rhythms in the thalamus of intact-cortex and decorticated cats. Journal of Neurophysiology 76: 4152–68.CrossRefGoogle ScholarPubMed
Timofeev, I. and Steriade, M. (1997) Fast (mainly 30–100 Hz) oscillations in the cat cerebellothalamic pathway and their synchronization with cortical potentials. Journal of Physiology (London) 504: 153–68.CrossRefGoogle ScholarPubMed
Timofeev, I. and Steriade, M. (1998) Cellular mechanisms underlying intrathalamic augmenting responses of reticular and relay neurons. Journal of Neurophysiology 79: 2716–29.CrossRefGoogle ScholarPubMed
Timofeev, I. and Steriade, M. (2004) Neocortical seizures: initiation, development and cessation. Neuroscience 123: 299–336.CrossRefGoogle ScholarPubMed
Timofeev, I., Contreras, D. and Steriade, M. (1996) Synaptic responsiveness of cortical and thalamic neurons during various phases of slow oscillation in cat. Journal of Physiology (London) 494: 265–78.CrossRefGoogle ScholarPubMed
Timofeev, I., Grenier, F., Bazhenov, M., Sejnowski, T. J. and Steriade, M. (2000) Origin of slow oscillations in deafferented cortical slabs. Cerebral Cortex 10: 1185–99.CrossRefGoogle ScholarPubMed
Timofeev, I., Bazhenov, M., Sejnowski, T. J. and Steriade, M. (2001a) Contribution of intrinsic and synaptic factors in the desynchronization of thalamic oscillatory activity. Thalamus and Related Systems 1: 53–69.Google Scholar
Timofeev, I., Grenier, F. and Steriade, M. (2001b) Disfacilitation and active inhibition in the neocortex during the natural sleep-wake cycle: an intracellular study. Proceedings of the National Academy of Sciences USA 98: 1924–9.CrossRefGoogle Scholar
Timofeev, I., Bazhenov, M., Sejnowski, T. J. and Steriade, M. (2002a).Cortical IH takes part in the generation of paroxysmal activities. Proceedings of the National Academy of Sciences USA 99: 9533–37.CrossRefGoogle Scholar
Timofeev, I., Grenier, F., Bazhenov, M., Houweling, A., Sejnowski, T. J. and Steriade, M. (2002b).Short- and medium-term plasticity associated with augmenting responses in cortical slabs and spindles in intact cortex of cats in vivo. Journal of Physiology (London) 542: 583–98.CrossRefGoogle Scholar
Tomberg, C. (1999) Cognitive N140 electrogenesis and concomitant 40 Hz synchronization in mid-dorsal prefrontal cortex (area 46) identified in non-averaged human brain potentials. Neuroscience Letters 266: 141–4.CrossRefGoogle Scholar
Tomberg, C. and Desmedt, J. E. (1999) The challenge of non-invasive cognitive physiology of the human brain: how to negotiate the irrelevant background noise without spoiling the recorded data through electronic averaging. Philosophical Transactions of the Royal Society of LondonB354: 1295–305.CrossRefGoogle ScholarPubMed
Tömböl, T. and Szafranska-Kosmal, A. (1972) A Golgi study of the amygdaloid complex in the cat. Acta Neurobiologiae Experimentalis 32: 835–48.Google ScholarPubMed
Tononi, G. and Cirelli, C. (2001) Some considerations on sleep and neural plasticity. Archives Italiennes de Biologie 139: 221–41.Google ScholarPubMed
Tononi, G. and Cirelli, C. (2003) Sleep and synaptic homeostasis. Brain Research Bulletin 62: 143–50.CrossRefGoogle ScholarPubMed
Toth, T. L. and Crunelli, V. (1992) Computer simulation of the pacemaker oscillations of thalamocortical cells. NeuroReport 3: 65–8.CrossRefGoogle ScholarPubMed
Toth, T. I., Hughes, S. W. and Crunelli, V. (1998) Analysis and biophysical interpretation of bistable behaviour in thalamocortical neurones. Neuroscience 87: 519–23.CrossRefGoogle Scholar
Traub, R. D., Miles, R. and Wong, R. K. S. (1989) Model of the origin of rhythmic population oscillations in the hippocampal slice. Science 243: 1319–25.CrossRefGoogle ScholarPubMed
Traub, R. D., Buhl, E. H., Glovell, T. and Whittington, M. A. (2003) Fast rhythmic bursting can be induced by in layer 2/3 cortical neurons by enhancing persistent Na+ conductance and blocking BK channels. Journal of Neurophysiology 89: 909–21.CrossRefGoogle ScholarPubMed
Traub, R. D., Contreras, D., Cunningham, M. O.et al. (2005) Single-column thalamocortical network model exhibiting gamma oscillations, sleep spindles, and epileptogenic bursts. Journal of Neurophysiology 93: 2194–232.CrossRefGoogle ScholarPubMed
Trulson, M. E. and Jacobs, B. L. (1979) Raphe unit activity in freely moving cats: correlation with level of behavioral arousal. Brain Research 163: 135–50.CrossRefGoogle ScholarPubMed
Tsatsanis, K. D., Rourke, B. P., Klin, A.et al. (2003) Reduced thalamic volume in high-functioning individuals with autism. Biological Psychiatry 53: 121–9.CrossRefGoogle ScholarPubMed
Tseng, K. Y., Kasanetz, F., Kargieman, L., Riquelne, L. A. and Murer, M. G. (2001) Cortical slow oscillatory activity is reflected in the membrane potential and spike trains of striatal neurons in rats with chronic nigrostriatal lesions. Journal of Neuroscience 21: 6430–9.CrossRefGoogle ScholarPubMed
Tsvetkov, E., Carlezon, W. Jr., Benes, F. M., Kandel, E. R. and Bolshakov, V. Y. (2002) Fear conditioning occludes LTP-induced presynaptic enhancement of synaptic transmission in the cortical pathway to the lateral amygdala. Neuron 34: 289–300.CrossRefGoogle ScholarPubMed
Tsvetkov, E., Shin, R. M. and Bolshakov, V. Y. (2004) Glutamate uptake determines pathway specificity of long-term potentiation in the neural circuitry of fear conditioning. Neuron 41: 139–51.CrossRefGoogle ScholarPubMed
Turner, B. H. and Herkenham, M. (1991) Thalamoamygdaloid projections in the rat: a test of the amygdala's role in sensory processing. Journal of Comparative Neurology 313: 295–325.CrossRefGoogle ScholarPubMed
Turner, B. H. and Zimmer, J. (1984) The architecture and some of the interconnections of the rat's amygdala and lateral periallocortex. Journal of Comparative Neurology 227: 540–57.CrossRefGoogle ScholarPubMed
Uchida, S., Maloney, T., March, J. D., Azari, R. and Feinberg, I. (1991) Sigma (12–15 Hz) and delta (0.3–3.0) Hz EEG oscillate reciprocally within NREM sleep. Brain Research Bulletin 27: 93–6.CrossRefGoogle Scholar
Ulrich, D. and Huguenard, J. R. (1996) GAB AB-receptor-mediated responses in GABAergic projection neurones of rat nucleus reticularis thalami in vitro. Journal of Physiology (London) 493: 845–54.CrossRefGoogle Scholar
Uhlrich, D. J., Manning, K. A. and Pienkowski, T. P. (1993) The histaminergic innervation of the lateral geniculate complex in the cat. Visual Neuroscience 10: 225–35.CrossRefGoogle ScholarPubMed
Uhlrich, D. J., Manning, K. A. and Xue, J. T. (2002) Effects of activation of the histaminergic tuberomammillary nucleus on visual responses of neurons in the dorsal lateral geniculate nucleus. Journal of Neuroscience 22: 1098–107.CrossRefGoogle ScholarPubMed
Umbriaco, D., Watkins, K. C., Descarries, L., Cozzari, C. and Hartman, B. K. (1994) Ultrastructural and morphometric features of the acetylcholine innervation in adult rat parietal cortex. An electron microscopic study in serial sections. Journal of Comparative Neurology 348: 351–73.CrossRefGoogle ScholarPubMed
Urbain, N., Gervasoni, D., Soulière, F.et al. (2000) Unrelated course of subthalamic nucleus and globus pallidus neuronal activities across vigilance states in the rat. European Journal of Neuroscience 12: 3361–74.CrossRefGoogle ScholarPubMed
Urbain, N., Rentéro, N., Gervasoni, D., Renaud, B. and Chouvet, G. (2002) The switch of subthalamic neurons from an irregular to a bursting pattern does not solely depend on their GABAergic inputs in the anesthetic-free rat. Journal of Neuroscience 22: 8665–75.CrossRefGoogle Scholar
Ursin, H. and Kaada, B. R. (1960) Functional localization within the amygdaloid complex in the cat. Electroencephalography and Clinical Neurophysiology 12: 1–20.CrossRefGoogle ScholarPubMed
Uva, L., Gruschke, S., Biella, G., Curtis, M. and Witter, M. P. (2004) Cytoarchitectonic characterization of the parahippocampal region in guinea pigs. Journal of Comparative Neurology 474: 289–303.CrossRefGoogle Scholar
Brederode, J. and Spain, W. (1995) Differences in inhibitory synaptic input between layer II–III and layer V neurons of the cat neocortex. Journal of Neurophysiology 74: 1149–66.CrossRefGoogle Scholar
Hoesen, G. W. and Pandya, D. N. (1975) Some connections of the entorhinal (area 28) and perirhinal (area 35) cortices of the rhesus monkey. I. Temporal lobe afferents. Brain Research 95: 1–24.CrossRefGoogle ScholarPubMed
Vanni-Mercier, G. and Debilly, G. (1998) A key role for the caudoventral pontine tegmentum in the simultaneous generation of eye saccades in bursts and associated ponto-geniculo-occipital waves during paradoxical sleep in the cat. Neuroscience 86: 571–85.CrossRefGoogle ScholarPubMed
Veening, J. G., Swanson, L. W. and Sawchenko, P. E. (1984) The organization of projections from the central nucleus of the amygdala to brainstem sites involved in central autonomic regulation: a combined retrograde transport-immunohistochemical study. Brain Research 303: 337–57.CrossRefGoogle ScholarPubMed
Velayos, J. L., Jimenez-Castellanos, J. Jr. and Reinoso-Suárez, F. (1989) Topographical organization of the projections from the reticular thalamic nucleus to the intralaminar and medial thalamic nuclei in the cat. Journal of Comparative Neurology 279: 457–69.CrossRefGoogle ScholarPubMed
Venance, L., Rozov, A., Blatow, M., Burnashev, N., Feldmeyer, D. and Monyer, H. (2000) Connexin expression in electrically coupled postnatal rat brain neurons. Proceedings of the National Academy of Sciences USA 97: 10260–5.CrossRefGoogle ScholarPubMed
Vertes, R. P. (1991) A PHA-L analysis of ascending projections of the dorsal raphe nucleus in the rat. Journal of Comparative Neurology 313: 643–68.CrossRefGoogle ScholarPubMed
Vertes, R. P. (2004) Differential projections of the infralimbic and prelimbic cortex in the rat. Synapse 51: 32–58.CrossRefGoogle ScholarPubMed
Vertes, R. P., Fortin, W. J. and Crane, A. M. (1999) Projections of the median raphe nucleus in the rat. Journal of Comparative Neurology 407: 555–82.3.0.CO;2-E>CrossRefGoogle ScholarPubMed
Villablanca, J. (1965) The electrocorticogram in the chronic cerveau isolé cat. Electroencephalography and Clinical Neurophysiology 19: 576–86.CrossRefGoogle ScholarPubMed
Villablanca, J. (1974) Role of the thalamus in sleep control: sleep-wakefulness studies of chronic cats without the thalamus: the ‘athalamic cat’. In Basic Sleep Mechanisms, ed. Petre-Quadens, O. and Schlag, J., pp. 51–81. New York: Academic Press.Google Scholar
Vincent, S. R. and Reiner, P. B. (1987) The immunohistochemical localization of choline acetyltransferase in the cat brain. Brain Research Bulletin 18: 371–415.CrossRefGoogle ScholarPubMed
Vincent, S. R., Satoh, K., Armstrong, D. M. and Fibiger, H. C. (1983) Substance P in the ascending cholinergic reticular system. Nature 306: 688–91.CrossRefGoogle ScholarPubMed
Vincent, S. R., Satoh, K., Armstrong, D. M.et al. (1986) Neuropeptides and NADPH-diaphorase activity in the ascending cholinergic reticular system of the rat. Neuroscience 17: 167–82.CrossRefGoogle ScholarPubMed
der Malsburg, C. (1995) Binding in models of perception and brain function. Current Opinions in Neurobiology 5: 520–6.CrossRefGoogle Scholar
der Malsburg, C. (1999) The what and why of binding: the modeler's perspective. Neuron 24: 95–104.CrossRefGoogle Scholar
Krosigk, M., Bal, T. and McCormick, D. A. (1993) Cellular mechanisms of a synchronized oscillation in the thalamus. Science 261: 361–4.CrossRefGoogle Scholar
Wainer, B. H. and Mesulam, M.-M. (1990) Ascending cholinergic pathways in the rat brain. In Brain Cholinergic Systems, ed. Steriade, M. and Biesold, D., pp. 65–119. Oxford: Oxford University Press.Google Scholar
Wainer, B. H., Bolam, J. P., Freund, T. F.et al. (1984) Cholinergic synapses in the rat brain: a correlated light and electron microscopic immunohistochemical study employing a monoclonal antibody against choline acetyltransferase. Brain Research 308: 69–76.CrossRefGoogle ScholarPubMed
Walker, D. L. and Davis, M. (2000) Involvement of NMDA receptors within the amygdala in short- versus long-term memory for fear conditioning as assessed with fear-potentiated startle. Behavioral Neuroscience 114: 1019–33.CrossRefGoogle ScholarPubMed
Walker, D. L. and Davis, M. (2002) The role of amygdala glutamate receptors in fear learning, fear-potentiated startle, and extinction. Pharmacology Biochemistry and Behavior 71: 379–92.CrossRefGoogle ScholarPubMed
Walker, D. L., Ressler, K. J., Lu, K. T. and Davis, M. (2002) Facilitation of conditioned fear extinction by systemic administration or intra-amygdala infusions of D-Cycloserine as assessed with fear-potentiated startle in rats. Journal of Neuroscience 22: 2343–51.CrossRefGoogle ScholarPubMed
Wallace, D. M., Magnuson, D. J. and Gray, T. S. (1989) The amygdalo-brainstem pathway: selective innervation of dopaminergic, noradrenergic and adrenergic cells in the rat. Neuroscience Letters 97: 252–8.CrossRefGoogle ScholarPubMed
Wallace, D. M., Magnuson, D. J. and Gray, T. S. (1992) Organization of amygdaloid projections to brainstem dopaminergic, noradrenergic, and adrenergic cell groups in the rat. Brain Research Bulletin 28: 447–54.CrossRefGoogle ScholarPubMed
Wan, H., Warburton, E. C., Zhu, X. O.et al. (2004) Benzodiazepine impairment of perirhinal cortical plasticity and recognition memory. European Journal of Neuroscience 20: 2214–24.CrossRefGoogle ScholarPubMed
Wang, X. J. (1999) Fast burst firing and short-term synaptic plasticity: a model of neocortical chattering neurons. Neuroscience 89: 347–62.CrossRefGoogle ScholarPubMed
Wang, X. J. and Rinzel, J. (1993) Spindle rhythmicity in the reticularis thalami nucleus: synchronization among mutually inhibitory neurons. Neuroscience 53: 899–904.CrossRefGoogle ScholarPubMed
Wang, X. J., Tegnér, J., Constantinidis, C. and Goldman-Rakic, P. S. (2004) Division of labor among distinct subtypes of inhibitory neurons in a cortical microcircuit of working memory. Proceedings of the National Academy of Sciences USA 101: 1368–73.CrossRefGoogle Scholar
Wang, Z. and McCormick, D. A. (1993) Control of firing mode of corticotectal and corticopontine layer V burst-generating neurons by norepinephrine, acetylcholine and 1S, 3R-ACPD. Journal of Neuroscience 13: 2199–216.CrossRefGoogle Scholar
Warburton, E. C., Koder, T., Cho, K.et al. (2003) Cholinergic neurotransmission is essential for perirhinal cortical. Neuron 38: 987–96.CrossRefGoogle ScholarPubMed
Washburn, M. S. and Moises, H. C. (1992a) Electrophysiological and morphological properties of rat basolateral amygdaloid neurons in vitro. Journal of Neuroscience 12: 4066–79.CrossRefGoogle Scholar
Washburn, M. S. and Moises, H. C. (1992b) Inhibitory responses of rat basolateral amygdaloid neurons recorded in vitro. Neuroscience 50: 811–30.CrossRefGoogle Scholar
Washburn, M. S. and Moises, H. C. (1992c) Muscarinic responses of rat basolateral amygdaloid neurons recorded in vitro. Journal of Physiology (London) 449: 121–54.CrossRefGoogle Scholar
Watanabe, T., Taguchi, Y., Shiosaka, S.et al. (1984) Distribution of the histaminergic neuron system in the central nervous system of rats; a fluorescent immunohistochemical analysis with histidine decarboxylase as a marker. Brain Research 295: 13–25.CrossRefGoogle ScholarPubMed
Webster, H. H. and Jones, B. E. (1988) Neurotoxic lesions of the dorsolateral pontomesencephalic tegmentum cholinergic area in the cat. II. Effects upon sleep-waking states. Brain Research 458: 285–302.CrossRefGoogle ScholarPubMed
Weese, G. D., Phillips, J. M. and Brown, V. J. (1999) Attentional orienting is impaired by unilateral lesions of the thalamic reticular nucleus in the rat. Journal of Neuroscience 19: 10135–9.CrossRefGoogle ScholarPubMed
Weisskopf, M. G., Bauer, E. P. and LeDoux, J. E. (1999) L-type voltage-gated calcium channels mediate NMDA-independent associative long-term potentiation at thalamic input synapses to the amygdala. Journal of Neuroscience 19: 10512–19.CrossRefGoogle ScholarPubMed
Weliky, M. and Katz, L. C. (1999) Correlational structure of spontaneous neuronal activity in the developing lateral geniculate nucleus in vivo. Science 285: 599–604.CrossRefGoogle ScholarPubMed
Westgaard, F. H., Bonato, P. and Holte, K. A. (2000) Low-frequency oscillations (<0.3 Hz) in the electromyographic (EMG) activity of the human trapezius muscle during sleep. Journal of Neurophysiology 88: 1177–84.CrossRefGoogle Scholar
Weyand, T. G., Boudreaux, M. and Guido, W. (2001) Burst and tonic response modes in thalamic neurons during sleep and wakefulness. Journal of Neurophysiology 85: 1107–18.CrossRefGoogle ScholarPubMed
Whalen, P. J., Rauch, S. L., Etcoff, N. L.et al. (1998) Masked presentations of emotional facial expressions modulate amygdala activity without explicit knowledge. Journal of Neuroscience 18: 411–8.CrossRefGoogle ScholarPubMed
White, E. L. (1989) Cortical Circuits: Synaptic Organization of the Cerebral Cortex. Boston, MA: Birkhäuser.CrossRefGoogle Scholar
Wiedermann, U. A. and Lüthi, A. (2003) Timing of network synchronization by refractory mechanisms. Journal of Neurophysiology 90: 3902–11.CrossRefGoogle Scholar
Wilensky, A. E., Schafe, G. E. and LeDoux, J. E. (1999) Functional inactivation of the amygdala before but not after auditory fear conditioning prevents memory formation. Journal of Neuroscience 19: RC48.CrossRefGoogle Scholar
Wilensky, A. E., Schafe, G. E. and LeDoux, J. E. (2000) Functional inactivation of amygdala nuclei during acquisition of Pavlovian fear conditioning. Society for Neuroscience Abstracts 26: 465.Google Scholar
Wilensky, A. E., Schafe, G. E. and LeDoux, J. E. (2001) Does the central nucleus of the amygdala contribute to the consolidation of auditory fear conditioning. Society for Neuroscience Abstracts 27: 187.Google Scholar
Williams, J. A. and Reiner, P. B. (1993) Noradrenaline hyperpolarizes identified rat mesopontine cholinergic neurons in vitro. Journal of Neuroscience 13: 3878–83.CrossRefGoogle ScholarPubMed
Williams, J. A., Comisarow, J., Day, J., Fibiger, H. C. and Reiner, P. B. (1994) State-dependent release of acetylcholine in the rat thalamus measured by in vivo microdialysis. Journal of Neuroscience 14: 5236–42.CrossRefGoogle ScholarPubMed
Williams, J. T., North, R. A., Shefner, S. A., Nishi, S. and Egan, T. M. (1984) Membrane properties of rat locus coeruleus neurones. Neuroscience 13: 137–56.CrossRefGoogle ScholarPubMed
Williams, S. M. and Goldman-Rakic, P. S. (1998) Widespread origin of the primate mesofrontal dopamine system. Cerebral Cortex 8: 321–45.CrossRefGoogle ScholarPubMed
Williams, S. R., Toth, T. I., Turner, J. P., Hughes, S. W. and Crunelli, V. (1997) The ‘window’ component of the low threshold Ca2+ current produces input signal amplification and bistability in cat and rat thalamocortical neurones. Journal of Physiology (London) 505: 689–705.CrossRefGoogle ScholarPubMed
Wilson, C. J. (1987) Morphology and synaptic connections of crossed corticostriatal neurons in the rat. Journal of Comparative Neurology 263: 567–80.CrossRefGoogle ScholarPubMed
Wilson, C. J. and Kawaguchi, Y. (1996) The origin of two-state spontaneous membrane potential fluctuations of neostriatal spiny neurons. Journal of Neuroscience 16: 2397–410.CrossRefGoogle Scholar
Wilson, F. A. and Rolls, E. T. (1990) Learning and memory is reflected in the response of reinforcement-related neurons in the primate basal forebrain. Journal of Neuroscience 10: 1254–67.CrossRefGoogle ScholarPubMed
Wilson, M. A. and McNaughton, B. L. (1994) Reactivation of hippocampal ensemble memories during sleep. Science 265: 676–9.CrossRefGoogle ScholarPubMed
Wilson, M. A., Mascagni, F. and McDonald, A. J. (2002) Sex differences in delta opioid receptor immunoreactivity in rat medial amygdala. Neuroscience Letters 328: 160–4.CrossRefGoogle ScholarPubMed
Witter, M. P. and Groenewegen, H. J. (1984) Laminar origin and septotemporal distribution of entorhinal and perirhinal projections to the hippocampus in the cat. Journal of Comparative Neurology 224: 371–85.CrossRefGoogle ScholarPubMed
Witter, M. P. and Groenewegen, H. J. (1986) Connections of the parahippocampal cortex in the cat. IV. Subcortical efferents. Journal of Comparative Neurology 251: 51–77.CrossRefGoogle Scholar
Witter, M. P., Room, P., Groenewegen, H. J. and Lohman, A. H. M. (1986) Connections of the parahippocampal cortex in the cat. V. Intrinsic connections: Comments on input/output connections with the hippocampus. Journal of Comparative Neurology 252: 78–94.CrossRefGoogle ScholarPubMed
Witter, M. P., Groenewegen, H. J., Lopes da Silva, F. H. and Lohman, A. H. M. (1989) Functional organization of the extrinsic and intrinsic circuitry of the parahippocampal region. Progress in Neurobiology 33: 161–253.CrossRefGoogle ScholarPubMed
Witter, M. P., Wouterlood, F. G., Naber, P. A. and Haeften, T. (2000) Anatomical organization of the parahippocampal-hippocampal network. Annals of the New York Academy of Sciences 911: 1–24.CrossRefGoogle ScholarPubMed
Woody, C. D., Gruen, E. and Wang, X. F. (2003) Electrical properties affecting discharges of units of the mid and posterolateral thalamus of conscious cats. Neuroscience 122: 531–9.CrossRefGoogle ScholarPubMed
Woolf, N. J., Eckenstein, F. and Butcher, L. L. (1984) Cholinergic systems in the rat brain: I. Projections to the limbic telencephalon. Brain Research Bulletin 13: 751–84.CrossRefGoogle ScholarPubMed
Woolf, N. J., Hernit, M. C. and Butcher, L. L. (1986) Cholinergic and non-cholinergic projections from the rat basal forebrain. Neuroscience Letters 66: 281–6.CrossRefGoogle ScholarPubMed
Wouterlood, F. G. (2002) Cell types, local connectivity, microcircuits, and distribution of markers. In The Parahippocampal Region, ed. Witter, M. P. and Wouterlood, F. G., pp. 61–84. Oxford: Oxford University Press.Google Scholar
Wouterlood, F., Mugnaini, E. and Nederlof, J. (1985) Projection of olfactory lobe efferents to layer I GABA-ergic neurons in the entorhinal area. Combination of anterograde degeneration and immunoelectron microscopy in rat. Brain Research 343: 283–96.CrossRefGoogle Scholar
Wouterlood, F. G., Härtig, W., Brückner, G. and Witter, M. P. (1995) Parvalbumin-immunoreactive neurons in the entorhinal cortex of the rat: Localization, morphology, connectivity and ultrastructure. Journal of Neurocytology 24: 135–53.CrossRefGoogle ScholarPubMed
Wouterlood, F. G., Van, D. J., Haeften, T. and Witter, M. P. (2000) Calretinin in the entorhinal cortex of the rat: distribution, morphology, ultrastructure of neurons, and colocalization with gamma-aminobutyric acid and parvalbumin. Journal of Comparative Neurology 425: 177–92.3.0.CO;2-G>CrossRefGoogle Scholar
Wu, J. Y., Guan, L. and Tsau, Y. (1999) Propagating activation during oscillations and evoked responses in neocortical slices. Journal of Neuroscience 19: 5005–15.CrossRefGoogle ScholarPubMed
Xiang, J. Z. and Brown, M. W. (1998) Differential neuronal encoding of novelty, familiarity and recency in regions of the anterior temporal lobe. Neuropharmacology 37: 657–76.CrossRefGoogle ScholarPubMed
Xiang, Z., Huguenard, J. R. and Prince, D. A. (1998) Cholinergic switching within neocortical inhibitory networks. Science 281: 985–8.CrossRefGoogle ScholarPubMed
Xu, L., Tripathy, A., Pasek, D. A. and Meissner, G. (1999) Ruthenium red modifies the cardiac and skeletal muscle Ca2+ release channels (ryanodine receptors) by multiple mechanisms. Journal of Biological Chemistry 274: 32680–91.CrossRefGoogle Scholar
Yajeya, J., Juan, A. D., Bajo, V. M.et al. (1999) Muscarinic activation of a non-selective cationic conductance in pyramidal neurons in rat basolateral amygdala. Neuroscience 88: 159–67.CrossRefGoogle ScholarPubMed
Yajeya, J., Fuente, A., Criado, J. M.et al. (2000) Muscarinic agonist carbachol depresses excitatory synaptic transmission in the rat basolateral amygdala in vitro. Synapse 38: 151–60.3.0.CO;2-K>CrossRefGoogle ScholarPubMed
Yamada, T., Kameyama, S., Fuchigami, Y.et al. (1988) Changes of short latency somatosensory evoked potentials in sleep. Electroencephalography and Clinical Neurophysiology 70: 126–36.CrossRefGoogle ScholarPubMed
Yang, C. R., Seamans, J. K. and Gorelova, N. (1996) Electrophysiological and morphological properties of layers V–VI principal pyramidal cells in rat prefrontal cortex in vitro. Journal of Neuroscience 16: 1904–21.CrossRefGoogle ScholarPubMed
Yen, C. T., Conley, M., Hendry, S. H. C. and Jones, E. G. (1985) The morphology of physiologically identified GABAergic neurons in the somatic sensory part of the thalamic reticular nucleus in the cat. Journal of Neuroscience 5: 2254–68.CrossRefGoogle ScholarPubMed
Yingling, C. D. and Skinner, J. E. (1977) Gating of thalamic input to cerebral cortex by nucleus reticularis thalami. In Attention, Voluntary Contraction and Event-Related Cerebral Potentials, ed. Desmedt, J., pp. 534–59. Basel: Karger.Google Scholar
Ylinen, A., Bragin, A., Nádasdy, Z.et al. (1995) Sharp wave-associated high-frequency oscillation (200 Hz) in the intact hippocampus: network and intracellular mechanisms. Journal of Neuroscience 15: 30–46.CrossRefGoogle Scholar
Young, M. P., Tanaka, K. and Yamane, S. (1992) On oscillating neuronal responses in the visual cortex of the monkey. Journal of Neurophysiology 67: 1464–74.CrossRefGoogle ScholarPubMed
Yu, Y. Q., Xiong, Y., Chan, Y. S. and He, J. (2004) Corticofugal gating of auditory information in the thalamus: an in vivo intracellular recording study. Journal of Neuroscience 24: 3060–9.CrossRefGoogle Scholar
Yuste, R. and Tank, D. W. (1996) Dendritic integration in mammalian neurons, a century after Cajal. Neuron 16: 701–16.CrossRefGoogle ScholarPubMed
Zhang, L. and Jones, E. G. (2004) Corticothalamic inhibition in the thalamic reticular nucleus. Journal of Neurophysiology 91: 759–66.CrossRefGoogle ScholarPubMed
Zhang, S. J., Huguenard, J. R. and Prince, D. A. (1997) GAB AA receptor-mediated Cl− currents in rat thalamic reticular and relay neurons. Journal of Neurophysiology 78: 2280–6.CrossRefGoogle Scholar
Zhu, J. J. and Heggelund, P. (2001) Muscarinic regulation of dendritic and axonal outputs of rat thalamic interneurons: a new cellular mechanism for uncoupling distal dendrites. Journal of Neuroscience 21: 1148–59.CrossRefGoogle ScholarPubMed
Zhu, J. J. and Lo, F. S. (1999) Three GABA receptor-mediated postsynaptic potentials in interneurons in the rat lateral geniculate nucleus. Journal of Neuroscience 19: 5721–30.CrossRefGoogle ScholarPubMed
Zhu, J. J. and Uhlrich, D. J. (1998) Cellular mechanisms underlying two muscarinic receptor-mediated depolarizing responses in relay cells of the rat lateral geniculate nucleus. Neuroscience 87: 767–81.CrossRefGoogle ScholarPubMed
Zhu, J. J., Lytton, W. W., Xue, J. T. and Uhlrich, D. J. (1999a) An intrinsic oscillation in interneurons of the rat lateral geniculate nucleus. Journal of Neurophysiology 81: 702–11.CrossRefGoogle Scholar
Zhu, J. J., Uhlrich, D. J. and Lytton, W. W. (1999b) Burst firing in identified rat geniculate interneurons. Neuroscience 91: 1445–60.CrossRefGoogle Scholar
Zhu, J. J., Uhlrich, D. J. and Lytton, W. W. (1999c) Properties of a hyperpolarization-activated cation current in interneurons in the rat lateral geniculate nucleus. Neuroscience 92: 445–57.CrossRefGoogle Scholar
Ziakopoulos, Z., Tillett, C. W., Brown, M. W. and Bashir, Z. I. (1999) Input- and layer-dependent synaptic plasticity in the rat perirhinal cortex in vitro. Neuroscience 92: 459–72.CrossRefGoogle ScholarPubMed
Zola-Morgan, S., Squire, L. R., Amaral, D. G. and Suzuki, W. A. (1989) Lesions of perirhinal and parahippocampal cortex that spare the amygdala and hippocampal formation produce severe memory impairment. Journal of Neuroscience 9: 4355–70.CrossRefGoogle ScholarPubMed
Abel, T., Nguyen, P. V., Barad, M.et al. (1997) Genetic demonstration of a role for PKA in the late phase of LTP and in hippocampus-based long-term memory. Cell 88: 615–26.CrossRefGoogle ScholarPubMed
Abeles, M., Bergman, H., Gat, I.et al. (1995) Cortical activity flips among quasi-stationary states. Proceedings of the National Academy of Sciences USA 92: 8616–20.CrossRefGoogle ScholarPubMed
Achermann, P. and Borbély, A. (1997) Low-frequency (<1 Hz) oscillations in the human sleep EEG. Neuroscience 81: 213–22.CrossRefGoogle Scholar
Adams, J. H., Graham, D. I. and Jennett, B. (2000) The neuropathology of the vegetative state after an acute brain insult. Brain 123: 1327–38.CrossRefGoogle ScholarPubMed
Adolphs, R., Cahill, L., Schul, R. and Babinsky, R. (1997) Impaired declarative memory for emotional stimuli following bilateral amygdala damage in humans. Learning and Memory 4: 291–300.CrossRefGoogle ScholarPubMed
Aggleton, J. P. (2000) The Amygdala: A Functional Analysis. Oxford: Oxford University Press.Google Scholar
Aggleton, J. P. and Mishkin, M. (1983a) Visual recognition impairment following medial thalamic lesions. Neuropsychologia 21: 189–97.CrossRefGoogle Scholar
Aggleton, J. P. and Mishkin, M. (1983b) Memory impairments following restricted medial thalamic lesions in monkeys. Experimental Brain Research 52: 199–209.CrossRefGoogle Scholar
Aggleton, J. P., Hunt, P. R. and Rawlins, J. N. (1986) The effects of hippocampal lesions upon spatial and non-spatial tests of working memory. Behavioral Brain Research 19: 133–46.CrossRefGoogle ScholarPubMed
Aghajanian, G. K. (1985) Modulation of a transient outward current in serotonergic neurones by alpha 1-adrenoceptors. Nature 315: 501–3.CrossRefGoogle ScholarPubMed
Aghajanian, G. K. and Wang, E. Y. (1977) Habenular and other midbrain raphe afferents demonstrated by a modified retrograde tracing technique. Brain Research 122: 229–42.CrossRefGoogle ScholarPubMed
Agranoff, B. W., Davis, R. E. and Brink, J. J. (1966) Chemical studies on memory fixation in goldfish. Brain Research 1: 303–9.CrossRefGoogle ScholarPubMed
Airaksinen, M. S. and Panula, P. (1988) The histaminergic system in the guinea pig central nervous system: an immunocytochemical mapping study using an antiserum against histamine. Journal of Comparative Neurology 273: 163–86.CrossRefGoogle ScholarPubMed
Aitkin, L. M., Irvine, D. R., Nelson, J. E., Merzenich, M. M. and Clarey, J. C. (1986) Frequency representation in the auditory midbrain and forebrain of a marsupial, the northern native cat(Dasyurus hallucatus). Brain Behavior and Evolution 29: 17–28.CrossRefGoogle Scholar
Alden, M., Besson, J. M. and Bernard, J. F. (1994) Organization of the efferent projections from the pontine parabrachial area to the bed nucleus of the stria terminalis and neighboring regions: a PHA-L study in the rat. Journal of Comparative Neurology 341: 289–314.CrossRefGoogle ScholarPubMed
Aleksanov, S. N. (1983) Coherent functions of the electrical activity of the hippocampus, amygdala and frontal cortex during alimentary instrumental reflexes in the dog. Zhurnal Vysshei Nervnoi Deyatelnosti Imeni I P Pavlova 33: 694–9.Google ScholarPubMed
Allison, T., McCarthy, G., Wood, C. C., Williamson, P. D. and Spencer, D. D. (1989) Human cortical potentials evoked by stimulation of the median nerve. II. Cytoarchitectonic areas generating long-latency activity. Journal of Neurophysiology 62: 711–22.CrossRefGoogle ScholarPubMed
Alloway, K. D., Wallace, M. B. and Johnson, M. J. (1994) Cross-correlation analysis of cuneothalamic interactions in the rat somatosensory system: influence of receptive field topography and comparisons with thalamocortical interactions. Journal of Neurophysiology 72: 1949–72.CrossRefGoogle ScholarPubMed
Alonso, A. (2002) Electrophysiology of neurones in the perirhinal and entorhinal cortices and neuromodulatory changes in firing patterns. In The Parahippocampal Region, ed. Witter, M. P. and Wouterlood, F., pp. 89–105. Oxford: Oxford University Press.Google Scholar
Alonso, A. and Garcia-Austt, E. (1987a) Neuronal sources of theta rhythm in the enthorinal cortex of the rat. I. Laminar distribution of theta field potentials. Experimental Brain Research 67: 493–501.CrossRefGoogle Scholar
Alonso, A. and Garcia-Austt, E. (1987b) Neuronal sources of theta rhythm in the enthorinal cortex of the rat. II. Phase relations between unit discharges and theta field potentials. Experimental Brain Research 67: 502–9.CrossRefGoogle Scholar
Alonso, A. and Klink, R. (1993) Differential electroresponsiveness of stellate and pyramidal-like cells of medial entorhinal cortex layer II. Journal of Neurophysiology 70: 128–43.CrossRefGoogle ScholarPubMed
Alonso, A. and Köhler, C. (1984) A study of the reciprocal connections between the septum and the entorhinal area using anterograde and retrograde axonal transport methods in the rat brain. Journal of Comparative Neurology 225: 327–43.CrossRefGoogle ScholarPubMed
Alonso, A. and Llinás, R. (1989) Subthreshold theta-like rhythmicity in stellate cells of enthorinal cortex layer II. Nature 342: 175–7.CrossRefGoogle Scholar
Alonso, A., Faure, M. F. and Beaudet, A. (1994) Neurotensin promotes oscillatory bursting behaviour and is internalized in basal forebrain cholinergic neurons. Journal of Neuroscience 14: 5778–92.CrossRefGoogle ScholarPubMed
Alonso, A., Khateb, A., Fort, P., Jones, B. E. and Mühlethaler, M. (1996) Differential oscillatory properties of cholinergic and noncholinergic nucleus basalis neurons in guinea pig brain slices. European Journal of Neuroscience 8: 169–82.CrossRefGoogle Scholar
Amaral, D. G. (1986) Amygdalohippocampal and amygdalocortical projections in the primate brain. Advances in Experimental Medicine and Biology 203: 3–17.CrossRefGoogle ScholarPubMed
Amaral, D. G. and Cowan, W. M. (1980) Subcortical afferents to the hippocampal formation in the monkey. Journal of Comparative Neurology 189: 573–91.CrossRefGoogle ScholarPubMed
Amaral, D. G. and Price, J. L. (1984) Amygdalo-cortical projections in the monkey(Macaca fascicularis). Journal of Comparative Neurology 230: 465–96.CrossRefGoogle Scholar
Amaral, D. G., Price, J. L., Pitkänen, A. and Carmichael, S. T. (1992) Anatomical organization of the primate amygdaloid complex. In The Amygdala: Neurobiological Aspects of Emotion, Memory, and Mental Dysfunction, ed. Aggleton, J. P., pp. 1–66. New York: Wiley-Liss.Google Scholar
Amorapanth, P., LeDoux, J. E. and Nader, K. (2000) Different lateral amygdala outputs mediate reactions and actions elicited by a fear-arousing stimulus. Nature Neuroscience 3: 74–9.CrossRefGoogle ScholarPubMed
Amzica, F. and Massimini, M. (2002) Glial and neuronal interactions during slow waves and paroxysmal activities in the neocortex. Cerebral Cortex 12: 1101–13.CrossRefGoogle Scholar
Amzica, F. and Neckelmann, D. (1999) Membrane capacitance of cortical neurons and glia during sleep oscillations and spike-wave seizures. Journal of Neurophysiology 82: 2731–46.CrossRefGoogle ScholarPubMed
Amzica, F. and Steriade, M. (1995a) Disconnection of intracortical synaptic linkages disrupts synchronization of a slow oscillation. Journal of Neuroscience 15: 4658–77.CrossRefGoogle Scholar
Amzica, F. and Steriade, M. (1995b) Short- and long-range neuronal synchronization of the slow (<1 Hz) cortical oscillation. Journal of Neurophysiology 75: 20–38.CrossRefGoogle Scholar
Amzica, F. and Steriade, M. (1996) Progressive cortical synchronization of ponto-geniculo-occipital potentials during rapid eye movement sleep. Neuroscience 72: 309–14.CrossRefGoogle ScholarPubMed
Amzica, F. and Steriade, M. (1997) The K-complex: its slow (<1 Hz) rhythmicity and relation to delta waves. Neurology 49: 952–9.CrossRefGoogle ScholarPubMed
Amzica, F. and Steriade, M. (1998a) Cellular substrates and laminar profile of sleep K-complex. Neuroscience 82: 671–86.CrossRefGoogle Scholar
Amzica, F. and Steriade, M. (1998b) Electrophysiological correlates of sleep delta waves. Electroencephalography and Clinical Neurophysiology 107: 69–83.CrossRefGoogle Scholar
Amzica, F. and Steriade, M. (2000) Neuronal and glial membrane potentials during sleep and paroxysmal oscillations in the cortex. Journal of Neuroscience 20: 6646–65.CrossRefGoogle Scholar
Amzica, F. and Steriade, M. (2002) The functional significance of K-complexes. Sleep Medicine Reviews 6: 139–49.CrossRefGoogle ScholarPubMed
Amzica, F., Neckelmann, D. and Steriade, M. (1997) Instrumental conditioning of fast (20- to 50 Hz) oscillations in corticothalamic networks. Proceedings of the National Academy of Sciences USA 94: 1985–9.CrossRefGoogle ScholarPubMed
Amzica, F., Massimini, M. and Manfridi, A. (2002) Spatial buffering during slow and paroxysmal oscillations in cortical networks of glial cellsin vivo. Journal of Neuroscience 22: 1042–53.CrossRefGoogle ScholarPubMed
Andersen, P. and Andersson, S. A. (1968) Physiological Basis of Alpha Rhythm. New York: Appleton-Century-Crofts.Google Scholar
Anderson, A. K. and Phelps, E. A. (2001) Lesions of the human amygdala impair enhanced perception of emotionally salient events. Nature 411: 305–9.CrossRefGoogle ScholarPubMed
Andreasen, N. C., Arndt, S., Swayze, V.et al. (1994) Thalamic abnormalities in schizophrenia visualied through magnetic resonance image averaging. Science 266: 294–8.CrossRefGoogle Scholar
Arieli, A., Shoham, D., Hildesheim, R. and Grinvald, A. (1995) Coherent spatiotemporal patterns of ongoing activity revealed by real-time optical imaging coupled with single-unit recording in the cat visual cortex. Journal of Neurophysiology 73: 2072–93.CrossRefGoogle ScholarPubMed
Arieli, A., Sterkin, A., Grinvald, A. and Aertsen, A. (1996) Dynamics of ongoing activity: explanation of the large variability in evoked cortical responses. Science 273: 1868–71.CrossRefGoogle ScholarPubMed
Asanuma, C. (1992) Noradrenergic innervation of the thalamic reticular nucleus: a light and electron microscopic immunohistochemical study in rats. Journal of Comparative Neurology 319: 299–311.CrossRefGoogle ScholarPubMed
Asanuma, C. (1997) Distribution of neuromodulatory inputs in the reticular and dorsal thalamic nuclei. In Thalamus, vol. 2 (Experimental and Clinical Aspects), ed. Steriade, M., Jones, E. G. and McCormick, D. A., pp. 93–153. Oxford: Elsevier.Google Scholar
Asanuma, C. and Porter, L. L. (1990) Light and electron microscopic evidence for a GABAergic projection from the caudal basal forebrian to the thalamic reticular nucleus in rats. Journal of Comparative Neurology 302: 159–72.CrossRefGoogle ScholarPubMed
Aston-Jones, G., Ennis, M., Pieribone, V. A., Nickell, W. T. and Shipley, M. T. (1986) The brain nucleus locus coeruleus: Restricted afferent control of a broad efferent network. Science 234: 734–7.CrossRefGoogle ScholarPubMed
Bacci, A., Huguenard, J. R. and Prince, D. A. (2003) Functional autaptic neurotransmission in fast-spiking interneurons: a novel form of feedback inhibition in the neocortex. Journal of Neuroscience 23: 859–66.CrossRefGoogle ScholarPubMed
Bahar, A., Samuel, A., Hazvi, S. and Dudai, Y. (2003) The amygdalar circuit that acquires taste aversion memory differs from the circuit that extinguishes it. European Journal of Neuroscience 17: 1527–30.CrossRefGoogle Scholar
Bailey, D. J., Kim, J. J., Sun, W., Thompson, R. F. and Helmstetter, F. J. (1999) Acquisition of fear conditioning in rats requires the synthesis of mRNA in the amygdala. Behavioral Neuroscience 113: 276–82.CrossRefGoogle ScholarPubMed
Baker, S. N., Olivier, E. and Lemon, R. N. (1997) Coherent oscillations in the monkey motor cortex and hand muscle EMG show task-dependent modulation. Journal of Physiology (London) 501: 225–41.CrossRefGoogle ScholarPubMed
Bal, T. and McCormick, D. A. (1996) What stops synchronized thalamocortical oscillations?Neuron 17: 297–308.CrossRefGoogle ScholarPubMed
Bal, T., Krosigk, M. and McCormick, D. A. (1995a) Synaptic and membrane mechanisms underlying synchronized oscillations in the ferret lateral geniculate nucleusin vitro. Journal of Physiology (London) 483: 641–63.CrossRefGoogle Scholar
Bal, T., Krosigk, M. and McCormick, D. A. (1995b) Role of the ferret perigeniculate nucleus in the generation of synchronized oscillationsin vitro. Journal of Physiology (London) 483: 665–85.CrossRefGoogle Scholar
Ball, G. J., Gloor, P. and Schaul, N. (1977) The cortical electromicrophysiology of pathological delta waves in the electroencephalogram of cats. Electroencephalography and Clinical Neurophysiology 43: 346–61.CrossRefGoogle ScholarPubMed
Barazangi, N. and Role, L. W. (2001) Nicotine-induced enhancement of glutamatergic and GABAergic synaptic transmission in the mouse amygdala. Journal of Neurophysiology 86: 463–74.CrossRefGoogle ScholarPubMed
Barbas, H. and Olmos, D, J. S. (1990) Projections from the amygdala to basoventral and mediodorsal prefrontal regions in the rhesus monkey. Journal of Comparative Neurology 300: 549–71.CrossRefGoogle ScholarPubMed
Bargas, J., Galarraga, E. and Aceves, J. (1989) An early outward conductance modulates the firing latency and frequency of neostriatal neurons of the rat brain. Experimental Brain Research 75: 146–56.CrossRefGoogle ScholarPubMed
Barria, A., Muller, D., Derkach, V., Griffith, L. C. and Soderling, T. R. (1997) Regulatory phosphorylation of AMPA-type glutamate receptors by CaM-KII during long-term potentiation. Science 276: 2042–5.CrossRefGoogle ScholarPubMed
Batsel, H. L. (1964) Spontaneous desynchronization in the chronic cat ‘cerveau isolé’. Archives Italiennes de Biologie 102: 547–66.Google Scholar
Battaglia, F. P., Sutherland, G. R. and McNaughton, B. L. (2004) Hippocampal sharp wave bursts coincide with neocortical ‘up-state’ transitions. Learning and Memory 11: 697–704.CrossRefGoogle ScholarPubMed
Bauer, E. P. and LeDoux, J. E. (2004) Heterosynaptic long-term potentiation of inhibitory interneurons in the lateral amygdala. Journal of Neuroscience 24: 9507–12.CrossRefGoogle ScholarPubMed
Bauer, E. P., LeDoux, J. E. and Nader, K. (2001) Fear conditioning and LTP in the lateral amygdala are sensitive to the same stimulus contingencies. Nature Neuroscience 4: 687–8.CrossRefGoogle ScholarPubMed
Baughman, R. W. and Gilbert, C. D. (1980) Aspartate and glutamate as possible neurotransmitters of cells in layer 6 of the visual cortex. Nature 287: 848–50.CrossRefGoogle ScholarPubMed
Baxter, M. G. and Murray, E. A. (2002) The amygdala and reward. Nature Reviews Neuroscience 3: 563–73.CrossRefGoogle ScholarPubMed
Bayer, L., Eggermann, E., Saint-Mieux, B.et al. (2002) Selective action of orexin (hypocretin) on nonspecific thalamocortical projection neurons. Journal of Neuroscience 22: 7835–9.CrossRefGoogle ScholarPubMed
Bayer, L., Serafin, M., Eggermann, E.et al. (2004) Exclusive postsynaptic action of hypocretin-orexin on sublayer 6b cortical neurons. Journal of Neuroscience 24: 6760–4.CrossRefGoogle ScholarPubMed
Bazhenov, M., Timofeev, I., Steriade, M. and Sejnowski, T. J. (1998a) Cellular and network models for intrathalamic augmenting responses during 10-Hz stimulation. Journal of Neurophysiology 79: 2730–48.CrossRefGoogle Scholar
Bazhenov, M., Timofeev, I., Steriade, M. and Sejnowski, T. J. (1998b) Computational models of thalamocortical augmenting responses. Journal of Neuroscience 18: 6444–65.CrossRefGoogle Scholar
Bazhenov, M., Timofeev, I., Steriade, M. and Sejnowski, T. J. (1999) Self-sustained rhythmic activity in the thalamic reticular nucleus mediated by depolarizing GABAA receptor potentials. Nature Neuroscience 2: 168–74.CrossRefGoogle ScholarPubMed
Bazhenov, M., Timofeev, I., Steriade, M. and Sejnowski, T. (2000) Spiking-bursting activity in the thalamic reticular nucleus initiates sequences of spindle oscillations in thalamic networks. Journal of Neurophysiology 84: 1076–87.CrossRefGoogle ScholarPubMed
Bazhenov, M., Timofeev, I., Steriade, M. and Sejnowski, T. J. (2002) Model of thalamocortical slow-wave sleep oscillations and transitions to activated states. Journal of Neuroscience 22: 8691–704.CrossRefGoogle ScholarPubMed
Bear, M. F. and Abraham, W. C. (1996) Long-term depression in hippocampus. Annual Review of Neuroscience 19: 437–62.CrossRefGoogle ScholarPubMed
Beaudet, A., and Descarries, L. (1978) The monoamine innervation of rat cerebral cortex: synaptic and non-synaptic axon terminals. Neuroscience 3: 851–60.CrossRefGoogle Scholar
Bechara, A., Tranel, D., Damasio, H.et al. (1995) Double dissociation of conditioning and declarative knowledge relative to the amygdala and hippocampus in humans. Science 269: 1115–18.CrossRefGoogle ScholarPubMed
Beckstead, R. M. (1978) Afferent connections of the entorhinal area in the rat as demonstrated by retrograde cell-labeling with horse radish peroxidase. Brain Research 152: 249–64.CrossRefGoogle Scholar
Beggs, J. M. and Kairiss, E. W. (1994) Electrophysiology and morphology of neurons in rat perirhinal cortex. Brain Research 665: 18–32.CrossRefGoogle ScholarPubMed
Behrendt, R. P. (2003) Hallucinations: synchronization of thalamocortical gamma oscillations underconstrained by sensory input. Consciousness and Cognition 12: 413–51.CrossRefGoogle Scholar
Bellgowan, P. S. and Helmstetter, F. J. (1996) Neural systems for the expression of hypoalgesia during nonassociative fear. Behavioral Neuroscience 110: 727–36.CrossRefGoogle ScholarPubMed
Ben-Ari, Y., Gal La Salle, G., Barbin, G., Schwartz, J. C. and Garbarg, M. (1977) Histamine synthesizing afferents within the amygdaloid complex and bed nucleus of the stria terminalis of the rat. Brain Research 138: 285–94.CrossRefGoogle ScholarPubMed
Bennett, M. V. L. (1970) Comparative physiology: electric organs. Annual Review of Physiology 32: 471–528.CrossRefGoogle ScholarPubMed
Bentivoglio, M., Aggleton, J. P. and Mishkin, M. (1997) The thalamus and memory formation. In Thalamus, vol. 2 (Experimental and Clinical Aspects), ed. Steriade, M., Jones, E. G. & McCormick, D. A., pp. 689–720. Oxford: Elsevier.
Berger, B., Thierry, A. M., Tassin, J. P. and Moyne, M. A. (1976) Dopaminergic innervation of the rat prefrontal cortex: a fluorescence histochemical study. Brain Research 106: 133–45.CrossRefGoogle ScholarPubMed
Berger, H. (1929) Uber das Elektroencephalogramm des Menschen. Archive für Psychiatrie und Nervenkrankheiten 87: 527–70.CrossRefGoogle Scholar
Berger, H. (1937) Uber das Elektroencephalogramm des Menschen. Dreizehnte Mitteilung. Archive für Psychiatrie und Nervenkrankheiten 106: 577–84.CrossRefGoogle Scholar
Berger, T., Senn, W. and Lüscher, H. R. (2003) Hyperpolarization-activated current Ih disconnects somatic and dendritic spike initiation zones in layer V pyramidal neurons. Journal of Neurophysiology 90: 2428–37.CrossRefGoogle ScholarPubMed
Berlucchi, G. (1997) One or many arousal systems? Reflections on some of Giuseppe Moruzzi's foresights and insights about intrinsic regulation of brain activity. Archives Italiennes de Biologie 135: 5–14.Google ScholarPubMed
Bernard, J. F., Huang, G. F. and Besson, J. M. (1990) Effect of noxious stimulation on the activity of neurons of the nucleus centralis of the amygdala. Brain Research 523: 347–50.CrossRefGoogle ScholarPubMed
Bernard, J. F., Huang, G. F. and Besson, J. M. (1992) Nucleus centralis of the amygdala and the globus pallidus ventralis: electrophysiological evidence for an involvement in pain processes. Journal of Neurophysiology 68: 551–69.CrossRefGoogle ScholarPubMed
Bernard, J. F., Alden, M. and Besson, J. M. (1993) The organization of the efferent projections from the pontine parabrachial area to the amygdaloid complex: A Phaseolus vulgaris leucoagglutinin (PHA-L) study in the rat. Journal of Comparative Neurology 329: 201–29.CrossRefGoogle ScholarPubMed
Berretta, S., Pantazopoulos, H., Caldera, M., Pantazopoulos, P. and Paré, D. (2005) Infralimbic cortex activation increases c-fos expression in intercalated neurons of the amygdala. Neuroscience 132: 943–53.CrossRefGoogle ScholarPubMed
Bickford, M. E., Günlük, A. E., Guido, W. and Sherman, S. M. (1993) Evidence that cholinergic axons from the parabrachial region of the brainstem are the exclusive source of nitric oxide in the lateral geniculate nucleus of the cat. Journal of Comparative Neurology 334: 410–30.CrossRefGoogle ScholarPubMed
Biella, G., Uva, L. and Curtis, M. (2001) Network activity evoked by neocortical stimulation in area 36 of the guinea pig perirhinal cortex. Journal of Neurophysiology 86: 164–72.CrossRefGoogle ScholarPubMed
Biella, G., Uva, L. and Curtis, M. (2002) Propagation of neuronal activity along the neocortical-perirhinal-entorhinal pathway in the guinea pig. Journal of Neuroscience 22: 9972–9.CrossRefGoogle ScholarPubMed
Bilkey, D. K. (1996) Long-term potentiation in the in vitro perirhinal cortex displays associative properties. Brain Research 733: 297–300.CrossRefGoogle ScholarPubMed
Bissière, S., Humeau, Y. and Lüthi, A. (2003) Dopamine gates LTP induction in lateral amygdala by suppressing feedforward inhibition. Nature Neuroscience 6: 587–92.CrossRefGoogle ScholarPubMed
Bland, B. H. (1986) The physiology and pharmacology of hippocampal formation theta rhythms. Progress in Neurobiology 26: 1–54.CrossRefGoogle ScholarPubMed
Bland, B. H., Andersen, P., Ganes, T. and Sven, O. (1980) Automated analysis of rhythmicity of physiologically identified hippocampal formation neurons. Experimental Brain Research 38: 205–19.CrossRefGoogle ScholarPubMed
Bliss, T. and Collingridge, G. L. (1993) A synaptic model of memory: long-term potentiation in the hippocampus. Nature 361: 31–9.CrossRefGoogle ScholarPubMed
Boejinga, P. H. and Lope, da Silva F. H. (1988) Differential distribution of beta and theta EEG activity in the entorhinal cortex of the cat. Brain Research 448: 272–86.CrossRefGoogle Scholar
Bordi, F., LeDoux, J., Clugnet, M. C. and Pavlides, C. (1993) Single-unit activity in the lateral nucleus of the amygdala and overlying areas of the striatum in freely behaving rats: rates, discharge patterns and responses to acoustic stimuli. Behavioral Neuroscience 107: 757–69.CrossRefGoogle ScholarPubMed
Bouyer, J. J., Montaron, M. F., Vahnée, J. M., Albert, M. P. and Rougeul, A. (1987) Anatomical localization of cortical beta rhythm in cat. Neuroscience 22: 863–9.CrossRefGoogle ScholarPubMed
Braga, M. F., Aroniadou-Anderjaska, V., Manion, S. T., Hough, C. J. and Li, H. (2004) Stress impairs alpha(1A) adrenoceptor-mediated noradrenergic facilitation of GABAergic transmission in the basolateral amygdala. Neuropsychopharmacology 29: 45–58.CrossRefGoogle ScholarPubMed
Bragin, A., Jandó, G., Nádasdy, Z.et al. (1995a) Gamma (40–100 Hz) oscillation in the hippocampus of the behaving rat. Journal of Neuroscience 15: 47–60.CrossRefGoogle Scholar
Bragin, A., Jandó, G., Nádasdy, Z., Van, L. M. and Buzsáki, G. (1995b) Dentate EEG spikes and associated interneuronal population bursts in the hippocampal hilar region of the rat. Journal of Neurophysiology 73: 1691–705.CrossRefGoogle Scholar
Braitenberg, V. and Schütz, A. (1991) Anatomy of the Cortex. Berlin: Springer.CrossRefGoogle Scholar
Braun, A. R., Balkin, T. J., Wesensten, N. J.et al. (1997) Regional cerebral blood flow throughout the sleep-wake cycle. Brain 120: 1173–97.CrossRefGoogle ScholarPubMed
Brazier, M. A. B. (1961) A History of the Electrical Activity of the Brain. London: Pitman.Google Scholar
Brazier, M. A. B. (1968) Studies of the EEG activity of limbic structures in man. Electroencephalography and Clinical Neurophysiology 25: 309–18.CrossRefGoogle ScholarPubMed
Bremer, F. (1935) Cerveau ‘isolé’ et physiologie du sommeil. Comptes Rendus de la Société de Biologie (Paris) 118: 1235–41.Google Scholar
Bremer, F. and Stoupel, N. (1959) Facilitation et inhibition des potentiels évoqués corticaux dans l'éveil cérébral. Archives Internationales de Physiologie 67: 240–75.CrossRefGoogle Scholar
Bremer, F. and Terzuolo, C. (1954) Contribution à l'étude des mécanismes physiologiques du maintien de l'activité vigile du cerveau. Interaction de la formation réticulée et de l'écorce cérébrale dans le processus de l'éveil. Archives Internationales de Physiologie 62: 157–78.Google Scholar
Bremer, F., Stoupel, N. and Reeth, P. C. (1960) Nouvelles recherches sur la facilitation et l'inhibition des potentiels évoqués corticaux dans l'éveil réticulaire. Archives Italiennes de Biologie 98: 229–47.Google Scholar
Bremner, J. D., Staib, L. H., Kaloupek, D.et al. (1999) Neural correlates of exposure to traumatic pictures and sound in Vietnam combat veterans with and without posttraumatic stress disorder: A positron emission tomography study. Biological Psychiatry 45: 806–16.CrossRefGoogle ScholarPubMed
Bringuier, V., Frégnac, Y., Baranyi, A., Debanne, D. and Shulz, D. E. (1997) Synaptic origin and stimulus dependency of neuronal oscillatory activity in the primary visual cortex of the cat. Journal of Physiology (London) 500: 751–74.CrossRefGoogle ScholarPubMed
Brinley-Reed, M., Mascagni, F. and McDonald, A. J. (1995) Synaptology of prefrontal projections to the basolateral amygdala: an electron microscopic study in the rat. Neuroscience Letters 202: 45–8.CrossRefGoogle ScholarPubMed
Brodmann, K. (1905) Beitraege zur histologischen Lokalisation der Grosshirnrinde. III Mitteilung. Die Rindenfelder der niederen Affen. Journal für Psychologie und Neurologie 4: 177–226.Google Scholar
Brodmann, K. (1908) Beitraege zur histologischen Lokalisation der Grosshirnrinde. VI Mitteilung. Die Cortexgliederung des Menschen. Journal für Psychologie und Neurologie 10: 231–46.Google Scholar
Brown, M. W. and Aggleton, J. P. (2001) Recognition memory: what are the roles of the perirhinal cortex and hippocampus?Nature Reviews Neuroscience 2: 51–61.CrossRefGoogle ScholarPubMed
Brown, M. W. and Bashir, Z. I. (2002) Evidence concerning how neurons of the perirhinal cortex may affect familiarity discrimination. Philosophical Transactions of the Royal Society of LondonB357: 1083–95.CrossRefGoogle Scholar
Brown, M. W., Wilson, F. A. W. and Riches, I. P. (1987) Neuronal evidence that inferomedial temporal cortex is more important than hippocampus in certain processes underlying recognition memory. Brain Research 409: 158–67.CrossRefGoogle ScholarPubMed
Brown, R. E., Stevens, D. R. and Haas, H. L. (2001) The physiology of brain histamine. Progress in Neurobiology 63: 637–72.CrossRefGoogle ScholarPubMed
Buchanan, S. L., Thompson, R. H., Maxwell, B. L. and Powell, D. A. (1994) Efferent connections of the medial prefrontal cortex in the rabbit. Experimental Brain Research 100: 469–83.CrossRefGoogle ScholarPubMed
Budde, T., Munsch, T. and Pape, H. C. (1998) Distribution of L-type calcium channels in rat thalamic neurones. European Journal of Neuroscience 10: 586–97.CrossRefGoogle ScholarPubMed
Burette, F., Jay, T. M. and Laroche, S. (1997) Reversal of LTP in the hippocampal afferent fiber system to the prefrontal cortex in vivo with low-frequency patterns of stimulation that do not produce LTD. Journal of Neurophysiology 78: 1155–60.CrossRefGoogle Scholar
Burlhis, T.M and Aghajanian, G. K. (1987) Pacemaker potentials of serotonergic dorsal raphe neurons: contribution of a low-threshold C a2+ conductance. Synapse 1: 582–8.CrossRefGoogle Scholar
Burwell, R. D. (2001) Borders and cytoarchitecture of the perirhinal and postrhinal cortices in the rat. Journal of Comparative Neurology 437: 17–41.CrossRefGoogle ScholarPubMed
Burwell, R. D. and Witter, M. P. (2002) Basic anatomy of the parahippocampal region in monkeys and rats. In The Parahippocampal Region, ed. Witter, M. P. and Wouterlood, F., pp. 35–59. New York: Oxford University Press.Google Scholar
Buzsáki, G. (1986) Hippocampal sharp waves: their origin and significance. Brain Research 398: 242–52.CrossRefGoogle ScholarPubMed
Buzsáki, G. (1989) Two-stage model of memory trace formation: a role for ‘noisy’ brain states. Neuroscience 31: 551–70.CrossRefGoogle ScholarPubMed
Buzsáki, G. (1996) The hippocampo-neocortical dialogue. Cerebral Cortex 6: 81–92.CrossRefGoogle ScholarPubMed
Buzsáki, G. (2002) Theta oscillations in the hippocampus. Neuron 33: 325–40.CrossRefGoogle ScholarPubMed
Buzsáki, G. and Draguhn, A. (2004) Neuronal oscillations in cortical networks. Science 304: 1926–9.CrossRefGoogle ScholarPubMed
Buzsáki, G. and Eidelberg, E. (1982) Direct afferent excitation and long-term potentiation of hippocampal interneurons. Journal of Neurophysiology 48: 597–607.CrossRefGoogle ScholarPubMed
Buzsáki, G., Leung, L. and Vanderwolf, C. H. (1983) Cellular bases of hippocampal EEG in the behaving rat. Brain Research 6: 139–71.CrossRefGoogle Scholar
Buzsáki, G., Czopf, J., Kondákor, I. and Kellényi, L. (1986) Laminar distribution of hippocampal rhythmic slow activity (RSA) in the behaving rat: current-source density analysis, effects of urethane and atropine. Brain Research 365: 125–37.CrossRefGoogle ScholarPubMed
Buzsáki, G., Bickford, R. G., Ponomareff, G.et al. (1988) Nucleus basalis and thalamic control of neocortical activity in the freely moving rat. Journal of Neuroscience 8: 4007–26.CrossRefGoogle ScholarPubMed
Buzsáki, G., Kennedy, B., Solt, V. B. and Ziegler, M. (1991) Noradrenergic control of thalamic oscillations: the role of α-2 receptors. European Journal of Neuroscience 3: 222–9.CrossRefGoogle Scholar
Buzsáki, G., Horváth, Z., Urioste, R., Hetke, J. and Wise, K. (1992) High-frequency network oscillation in the hippocampus. Science 256: 1025–7.CrossRefGoogle ScholarPubMed
Cahill, L. and McGaugh, J. L. (1998) Mechanisms of emotional arousal and lasting declarative memory. Trends in Neurosciences 21: 294–9.CrossRefGoogle ScholarPubMed
Cahill, L., Babinsky, R., Markowitsch, H. J. and McGaugh, J. L. (1995) The amygdala and emotional memory. Nature 377: 295–6.CrossRefGoogle ScholarPubMed
Cahill, L., Haier, R. J., Fallon, J.et al. (1996) Amygdala activity at encoding correlated with long-term, free recall of emotional information. Proceedings of the National Academy of Sciences USA 93: 8016–21.CrossRefGoogle ScholarPubMed
Cahill, L., Weinberger, N. M., Roozendaal, B. and McGaugh, J. L. (1999) Is the amygdala a locus of ‘conditioned fear’? Some questions and caveats. Neuron 23: 227–8.CrossRefGoogle Scholar
Cantero, J. L., Atienza, M., Salas, R. M. and Dominguez-Marin, E. (2002) Effects of prolonged waking-auditory stimulation on electroencephalogram synchronization and cortical coherence during subsequent slow-wave sleep. Journal of Neuroscience 22: 4702–8.CrossRefGoogle ScholarPubMed
Cape, E. G. and Jones, B. E. (2000) Effects of glutamate agonist versus procaine microinjections into the basal forebrain cholinergic cell area upon gamma and theta EEG activity and sleep-wake state. European Journal of Neuroscience 12: 2166–84.CrossRefGoogle ScholarPubMed
Cardin, J. A., Palmer, L. A. and Contreras, D. (2005) Stimulus-dependent gamma (30–50 Hz) oscillations in simple and complex fast rhythmic bursting cells in primary visual cortex. Journal of Neuroscience 25: 5339–50.CrossRefGoogle ScholarPubMed
Carlsen, J. (1988) Immunocytochemical localization of glutamate decarboxylase in the rat basolateral amygdaloid nucleus, with special reference to GABAergic innervation of amygdalostriatal projection neurons. Journal of Comparative Neurology 273: 513–26.CrossRefGoogle ScholarPubMed
Carlsen, J. and Heimer, L. (1986) A correlated light and electron microscopic immunocytochemical study of cholinergic terminals and neurons in the rat amygdaloid body with special emphasis on the basolateral amygdaloid nucleus. Journal of Comparative Neurology 244: 121–36.CrossRefGoogle ScholarPubMed
Carlsen, J., Zaborszky, L. and Heimer, L. (1985) Cholinergic projections from the basal forebrain to the basolateral amygdaloid complex: a combined retrograde fluorescent and immunohistochemical study. Journal of Comparative Neurology 234: 155–67.CrossRefGoogle ScholarPubMed
Carmichael, S. T. and Price, J. L. (1995a) Limbic connections of the orbital and medial prefrontal cortex in macaque monkeys. Journal of Comparative Neurology 363: 615–41.CrossRefGoogle Scholar
Carmichael, S. T. and Price, J. L. (1995b) Sensory and premotor connections of the orbital and medial prefrontal cortex of macaque monkeys. Journal of Comparative Neurology 363: 642–64.CrossRefGoogle Scholar
Carr, C. E. and Konishi, M. (1990) A circuit for detection of interaural time differences in the brainstem of the barn owl. Journal of Neuroscience 10: 3227–46.CrossRefGoogle ScholarPubMed
Carr, D. B. and Sesack, S. R. (2000a) Dopamine terminals synapse on callosal projection neurons in the rat prefrontal cortex. Journal of Comparative Neurology 425: 275–83.3.0.CO;2-Z>CrossRefGoogle Scholar
Carr, D. B. and Sesack, S. R. (2000b) GABA-containing neurons in the rat ventral tegmental area project to the prefrontal cortex. Synapse 38: 114–23.3.0.CO;2-R>CrossRefGoogle Scholar
Carr, D. B. and Sesack, S. R. (2000c) Projections from the rat prefrontal cortex to the ventral tegmental area: target specificity in the synaptic associations with mesoaccumbens and mesocortical neurons. Journal of Neuroscience 20: 3864–73.CrossRefGoogle Scholar
Carskadon, M. A. and Dement, W. C. (2000) Normal human sleep: an overview. In Principles and Practice of Sleep Medicine, ed. Kryger, M. H., Roth, T. and Dement, W. C., pp. 15–25. Philadelphia, PA: W. B. Saunders.Google Scholar
Cassell, M. D. and Gray, T. S. (1989a) Morphology of peptide-immunoreactive neurons in the rat central nucleus of the amygdala. Journal of Comparative Neurology 281: 320–33.CrossRefGoogle Scholar
Cassell, M. D. and Gray, T. S. (1989b) The amygdala directly innervates adrenergic (C1) neurons in the ventrolateral medulla in the rat. Neuroscience Letters 97: 163–8.CrossRefGoogle Scholar
Cassell, M. D., Gray, T. S. and Kiss, J. Z. (1986) Neuronal architecture in the rat central nucleus of the amygdala: a cytological, hodological, and immunocytochemical study. Journal of Comparative Neurology 246: 478–99.CrossRefGoogle ScholarPubMed
Castaigne, P., Buge, A., Escourolle, R. and Mason, M. (1962) Ramollissement pédonculaire médian, tegmentothalamique avec ophthalmoplégie et hypersomnie. Revue Neurologique (Paris) 106: 357–67.Google Scholar
Castellano, C., Brioni, J. D., Nagahara, A. H. and McGaugh, J. L. (1989) Post-training systemic and intra-amygdala administration of the GABA-B agonist baclofen impairs retention. Behavioral and Neural Biology 52: 170–9.CrossRefGoogle ScholarPubMed
Castelo-Branco, M., Neuenschwander, S. and Singer, W. (1998) Synchronization of visual responses between the cortex, lateral geniculate nucleus, and retina in the anesthetized cat. Journal of Neuroscience 18: 6395–410.CrossRefGoogle ScholarPubMed
Castro-Alamancos, M. A. (2002a) Properties of primary sensory (lemniscal) synapses in the ventrobasal thalamus and the relay of high-frequency sensory inputs. Journal of Neurophysiology 87: 946–53.CrossRefGoogle Scholar
Castro-Alamancos, M. A. (2002b) Different temporal processing of sensory inputs in the rat thalamus during quiescent and information processing states in vivo. Journal of Physiology (London) 539: 567–78.CrossRefGoogle Scholar
Castro-Alamancos, M. A. and Calcagnotto, M. E. (2001) High-pass filtering of corticothalamic activity by neuromodulators released in the thalamus during arousal: in vitro and in vivo. Journal of Neurophysiology 85: 1489–97.CrossRefGoogle ScholarPubMed
Castro-Alamancos, M. A. and Connors, B. W. (1996) Cellular mechanisms of the augmenting response: short-term plasticity in a thalamocortical pathway. Journal of Neuroscience 16: 7742–56.CrossRefGoogle Scholar
Cauli, B., Audinat, E., Lambolez, B.et al. (1997) Molecular and physiological diversity of cortical nonpyramidal cells. Journal of Neuroscience 17: 3894–906.CrossRefGoogle ScholarPubMed
Cauller, L. J. and Kulics, A. T. (1988) A comparison of awake and sleeping cortical states by analysis of the somatosensory-evoked response of postcentral area 1 in rhesus monkey. Experimental Brain Research 72: 584–92.CrossRefGoogle ScholarPubMed
Cavada, C. and Reinoso-Suarez, F. (1985) Topographical organization of the cortical afferent connections of the prefrontal cortex in the cat. Journal of Comparative Neurology 242: 293–324.CrossRefGoogle ScholarPubMed
Cederbaum, J. M. and Aghajanian, G. K. (1978) Afferent projections to the rat locus coeruleus as determined by a retrograde tracing technique. Journal of Comparative Neurology 178: 1–16.CrossRefGoogle Scholar
Chagnac-Amitai, Y. and Connors, B. W. (1989) Synchronized excitation and inhibition driven by bursting neurons in neocortex. Journal of Neurophysiology 62: 1149–62.CrossRefGoogle ScholarPubMed
Chapman, P. F. and Bellavance, L. L. (1992) Induction of long-term potentiation in the basolateral amygdala does not depend on NMDA receptor activation. Synapse 11: 310–18.CrossRefGoogle Scholar
Chapman, P. F., Kairiss, E. W., Keenan, C. L. and Brown, T. H. (1990) Long-term synaptic potentiation in the amygdala. Synapse 6: 271–8.CrossRefGoogle ScholarPubMed
Chen, J. C. and Lang, E. J. (2003) Inhibitory control of rat lateral amygdaloid projection cells. Neuroscience 121: 155–66.CrossRefGoogle ScholarPubMed
Chen, W., Zhang, J. J., Hu, G. Y. and Wu, C. P. (1996) Electrophysiological and morphological properties of pyramidal and non-pyramidal neurons in the cat motor cortex in vitro. Neuroscience 73: 39–55.CrossRefGoogle Scholar
Chi, C. C. and Flynn, J. P. (1971) Neuroanatomic projections related to biting attack elicited from hypothalamus in cats. Brain Research 35: 49–66.CrossRefGoogle ScholarPubMed
Cho, K., Kemp, N., Noel, J., Aggleton, J. P., Brown, M. W. and Bashir, Z. I. (2000) A new form of long-term depression in the perirhinal cortex. Nature Neuroscience 3: 150–6.CrossRefGoogle ScholarPubMed
Cho, K., Aggleton, J. P., Brown, M. W. and Bashir, Z. I. (2001) An experimental test of the role of postsynaptic calcium levels in determining synaptic strength using perirhinal cortex of rat. Journal of Physiology (London) 532: 459–66.CrossRefGoogle ScholarPubMed
Chow, A., Farb, C., Nadal, M. S.et al. (1999) K+ channel expression distinguishes subpopulations of parvalbumin- and somatostatin-containing neocortical interneurons. Journal of Neuroscience 19: 9332–45.CrossRefGoogle ScholarPubMed
Chrobak, J. J. and Buzsáki, G. (1994) Selective activation of deep layers (V–VI) retrohippocampal cortical neurons during hippocampal sharp waves in the behaving rat. Journal of Neuroscience 14: 6160–70.CrossRefGoogle ScholarPubMed
Chrobak, J. J. and Buzsáki, G. (1996) High-frequency oscillations in the output networks of the hippocampal-entorhinal axis of the freely behaving rat. Journal of Neuroscience 16: 3056–66.CrossRefGoogle ScholarPubMed
Chrobak, J. J. and Buzsáki, G. (1998) Gamma oscillations in the entorhinal cortex of the freely behaving rat. Journal of Neuroscience 18: 388–98.CrossRefGoogle ScholarPubMed
Cicogna, P, Natale, V., Occhionero, M and Bosinelli, M. (2000) Slow wave sleep and REM sleep mentation. Sleep Research Online 3: 67–72.Google ScholarPubMed
Cirelli, C. and Tononi, G. (2000) Differential expression of plasticity-related genes in waking and sleep and their regulation by the noradrenergic system. Journal of Neuroscience 20: 9184–7.CrossRefGoogle ScholarPubMed
Cirelli, C. and Tononi, G. (2004) Locus coeruleus control of state-dependent gene expression. Journal of Neuroscience 24: 5410–19.CrossRefGoogle ScholarPubMed
Cirelli, C., Pompeiano, M. and Tononi, G. (1996) Neuronal gene expression in the waking state: a role for the locus coeruleus. Science 274: 1211–15.CrossRefGoogle ScholarPubMed
Cirelli, C., Gutierrez, C. M. and Tononi, G. (2004) Extensive and divergent effects of sleep and wakefulness on brain gene expression. Neuron 41: 35–43.CrossRefGoogle ScholarPubMed
Cirelli, C., Huber, R., Gopalakrishnan, A., Southard, T. L. and Tononi, G. (2005) Locus coeruleus control of slow-wave homeostasis. Journal of Neuroscience 25: 4503–11.CrossRefGoogle ScholarPubMed
Cissé, Y., Grenier, F., Timofeev, I. and Steriade, M. (2003) Electrophysiological properties and input-output organization of callosal neurons in cat association cortex. Journal of Neurophysiology 89: 1402–13.CrossRefGoogle ScholarPubMed
Cissé, Y., Crochet, S., Timofeev, I. and Steriade, M. (2004) Synaptic enhancement induced through callosal pathways in cat association cortex. Journal of Neurophysiology 92: 3221–32.CrossRefGoogle ScholarPubMed
Clugnet, M. C. and LeDoux, J. E. (1990) Synaptic plasticity in fear conditioning circuits: induction of LTP in the lateral nucleus of the amygdala by stimulation of the medial geniculate body. Journal of Neuroscience 10: 2818–24.CrossRefGoogle ScholarPubMed
Coenen, A. M. L. and Vendrik, A. J. H. (1972) Determination of the transfer ratio of cat's geniculate neurons through quasi-intracellular recordings and the relation with the level of alertness. Experimental Brain Research 14: 227–42.CrossRefGoogle ScholarPubMed
Colino, A. and Fernánde, Molina, A. (1986) Electrical activity generated in the subicular complex and basolateral amygdala of the rat. Neuroscience 19: 573–80.CrossRefGoogle ScholarPubMed
Collins, D. R. and Paré, D. (2000) Differential fear conditioning induces reciprocal changes in the sensory responses of lateral amygdala neurons to the CS(+) and CS(−). Learning and Memory 7: 97–103.CrossRefGoogle Scholar
Collins, D. R., Lang, E. J. and Paré, D. (1999) Spontaneous activity of the perirhinal cortex in behaving cats. Neuroscience 89: 1025–39.CrossRefGoogle ScholarPubMed
Collins, D. R., Pelletier, J. G. and Paré, D. (2001) Slow and fast (gamma) neuronal oscillations in the perirhinal cortex and lateral amygdala. Journal of Neurophysiology 85: 1661–72.CrossRefGoogle ScholarPubMed
Colombo, M. and Gross, C. G. (1994) Responses of inferior temporal cortex and hippocampal neurons during delayed matching to sample in monkeys (Macaca fascicularis). Behavioral Neuroscience 108: 443–55.CrossRefGoogle Scholar
Colonnier, M. (1966) The structural design of the neocortex. In Brain and Conscious Experience, ed. Eccles, J. C., pp. 1–23. New York: Springer.Google Scholar
Compte, A., Sanchez-Vives, M. V., McCormick, D. A. and Wang, X. J. (2003) Cellular and network mechanisms of slow oscillatory activity (<1 Hz) and wave propagations in a cortical network model. Journal of Neurophysiology 89: 2707–25.CrossRefGoogle Scholar
Condé, F., Audinat, E., Maire-Lepoivre, E. and Crépel, F. (1990) Afferent connections of the medial frontal cortex of the rat. A study using retrograde transport of fluorescent dyes. I. Thalamic afferents. Brain Research Bulletin 24: 341–54.CrossRefGoogle Scholar
Condé, F., E., M.-L., Audinat, E. and Crépel, F. (1995) Afferent connections of the medial frontal cortex of the rat. II. Cortical and subcortical afferents. Journal of Comparative Neurology 352: 567–93.CrossRefGoogle ScholarPubMed
Condorelli, D. F., Belluard, N., Trovato-Salinaro, A. and Mudo, I. (2000) Expression of Cx36 in mammalian neurons. Brain Research Reviews 32: 72–85.CrossRefGoogle ScholarPubMed
Connors, B. W. and Amitai, Y. (1995) Functions of local circuits in neocortex: synchrony and laminae. In The Cortical Neuron, ed. Gutnick, M. J. and Mody, I., pp. 123–40. New York: Oxford University Press.Google Scholar
Connors, B. W. and Gutnick, M. J. (1990) Intrinsic firing patterns of diverse neocortical neurons. Trends in Neurosciences 13: 99–104.CrossRefGoogle ScholarPubMed
Connors, B. W., Gutnick, M. J. and Prince, D. A. (1982) Electrophysiological properties of neocortical neurons in vitro. Journal of Neurophysiology 48: 1302–20.CrossRefGoogle ScholarPubMed
Connors, B. W., Malenka, R. and Silva, L. R. (1988) Two inhibitory postsynaptic potentials, and GAB AA and GABAB receptor-mediated responses in neocortex of rat and cat. Journal of Physiology (London) 406: 443–68.CrossRefGoogle Scholar
Contreras, D. and Llinás, R. (2001) Voltage-sensitive dye imaging of neocortical spatio-temporal dynamics to afferent activation frequency. Journal of Neuroscience 21: 9403–13.CrossRefGoogle Scholar
Contreras, D. and Palmer, L. (2003) Response to contrast of electrophysiologically defined cell classes in primary visual cortex. Journal of Neuroscience 23: 6936–45.CrossRefGoogle ScholarPubMed
Contreras, D. and Steriade, M. (1995) Cellular basis of EEG slow rhythms: a study of dynamic corticothalamic relationships. Journal of Neuroscience 15: 604–22.CrossRefGoogle ScholarPubMed
Contreras, D. and Steriade, M. (1996) Spindle oscillation: the role of corticothalamic feedback in a thalamically generated rhythm. Journal of Physiology (London) 490: 159–79.CrossRefGoogle Scholar
Contreras, D. and Steriade, M. (1997a) Synchronization of low-frequency rhythms in corticothalamic networks. Neuroscience 76: 11–24.CrossRefGoogle Scholar
Contreras, D. and Steriade, M. (1997b) State-dependent fluctuations of low-frequency rhythms in corticothalamic networks. Neuroscience 76: 25–38.CrossRefGoogle Scholar
Contreras, D., Curró Dossi, R. and Steriade, M. (1993) Electrophysiological properties of cat reticular neurones in vivo. Journal of Physiology (London) 470: 273–94.CrossRefGoogle ScholarPubMed
Contreras, D., Destexhe, A., Sejnowski, T. J. and Steriade, M. (1996a) Control of spatiotemporal coherence of a thalamic oscillation by corticothalamic feedback. Science 274: 771–4.CrossRefGoogle Scholar
Contreras, D., Timofeev, I. and Steriade, M. (1996b) Mechanisms of long-lasting hyperpolarizations underlying slow sleep oscillations in cat corticothalamic networks. Journal of Physiology (London) 494: 251–64.CrossRefGoogle Scholar
Contreras, D., Destexhe, A. and Steriade, M. (1997a) Intracellular and computational characterization of the intracortical inhibitory control of synchronized thalamic inputs in vivo. Journal of Neurophysiology 78: 335–50.CrossRefGoogle Scholar
Contreras, D., Destexhe, A., Sejnowski, T. J. and Steriade, M. (1997b) Spatiotemporal patterns of spindle oscillations in cortex and thalamus. Journal of Neuroscience 17: 1179–96.CrossRefGoogle Scholar
Contreras, D., Dürmüller, N. and Steriade, M. (1997c) Absence of a prevalent laminar distribution of IPSPs in association cortical neurons of cat. Journal of Neurophysiology 78: 2742–53.CrossRefGoogle Scholar
Cossart, R., Aronov, D. and Yuste, R. (2003) Attractor dynamics of network UP states in the neocortex. Nature 423: 283–8.CrossRefGoogle ScholarPubMed
Cox, C. L. and Sherman, S. M. (2000) Control of dendritic outputs of inhibitory interneurons in the lateral geniculate nucleus. Neuron 27: 597–610.CrossRefGoogle ScholarPubMed
Cox, C. L., Huguenard, J. R. and Prince, D. A. (1996) Heterogenous axonal arborizations of rat thalamic reticular neurons in the ventrobasal nucleus. Journal of Comparative Neurology 366: 416–30.3.0.CO;2-7>CrossRefGoogle Scholar
Cox, C. L., Huguenard, J. R. and Prince, D. A. (1997) Nucleus reticularis neurons mediate diverse inhibitory effects in the thalamus. Proceedings of the National Academy of Sciences USA 94: 8854–9.CrossRefGoogle ScholarPubMed
Cox, C. L., Zhou, Q. and Sherman, S. M. (1998) Glutamate locally activates dendritic outputs of thalamic interneurons. Nature 394: 478–82.CrossRefGoogle ScholarPubMed
Cox, C. L., Reichova, I. and Sherman, S. M. (2003) Functional synaptic contacts by intranuclear axon collaterals of thalamic relay neurons. Journal of Neuroscience 23: 7642–6.CrossRefGoogle ScholarPubMed
Cox, G. E., Jordan, D., Moruzzi, P.et al. (1986) Amygdaloid influences on brain-stem neurones in the rabbit. Journal of Physiology (London) 381: 135–48.CrossRefGoogle ScholarPubMed
Crabtree, J. W., Collingridge, G. L. and Isaac, J. T. R. (1998) A new intrathalamic pathway linking modality-related nuclei in the dorsal thalamus. Nature Neuroscience 1: 389–94.CrossRefGoogle ScholarPubMed
Crick, F. (1984) The function of the thalamic reticular complex: the searchlight hypothesis. Proceedings of the National Academy of Sciences USA 81: 4586–90.CrossRefGoogle ScholarPubMed
Crick, F. (1994) The Astonishing Hypothesis. New York: Charles Scribner's Sons.Google Scholar
Crill, W. E. (1996) Persistent sodium current in mammalian central neurones. Annual Review of Physiology 58: 349–62.CrossRefGoogle Scholar
Crochet, S., Fuentealba, P., Timofeev, I. and Steriade, M. (2004) Selective amplification of neocortical neuronal output by fast prepotentialsin vivo. Cerebral Cortex 14: 1110–21.CrossRefGoogle ScholarPubMed
Crosby, E. C. and Humphrey, T. (1941) Studies of the vertebrate telencephalon. II. The nuclear pattern of the anterior olfactory nucleus, tuberculum olfactorum and the amygdaloid complex in adult man. Journal of Comparative Neurology 74: 309–52.CrossRefGoogle Scholar
Crow, T. J. (1997) Schizophrenia as failure of hemispheric dominance for language. Trends in Neuroscience 20: 339–43.Google ScholarPubMed
Crunelli, V. and Leresche, N. (2002) Childhood absence epilepsy: genes, channels, neurons and networks. Nature Reviews Neuroscience 3: 371–82.CrossRefGoogle ScholarPubMed
Crunelli, V., Haby, M., Jassik-Gerschenfeld, D., Leresche, N. and Pirchio, M. (1988) C l− and K+-dependent inhibitory postsynaptic potentials evoked by interneurons of the rat lateral geniculate nucleus. Journal of Physiology (London) 399: 153–76.CrossRefGoogle Scholar
Crutcher, M. D., Branch, M. H., DeLong, M. R. and Georgopoulos, A. P. (1980) Activity of zona incerta neurons in the behaving primate. Society for Neuroscience Abstracts 6: 676.Google Scholar
Csicsvari, J., Hirase, H., Czurkó, A., Mamiya, A. and Buzsáki, G. (1999) Oscillatory coupling of hippocampal pyramidal cells and interneurons in the behaving cat. Journal of Neuroscience 19: 274–87.CrossRefGoogle Scholar
Cunningham, E. T. and LeVay, S. (1986) Laminar and synaptic organization of the projection from the thalamic nucleus centralis to primary visual cortex in the cat. Journal of Comparative Neurology 254: 65–77.CrossRefGoogle ScholarPubMed
Cunningham, M. O., Whittington, M. A., Bibbig, A.et al. (2004) A role for fast rhythmic bursting neurons in cortical gamma oscillations in vitro. Proceedings of the National Academy of Sciences USA 101: 7152–7.CrossRefGoogle ScholarPubMed
Curró Dossi, R., Paré, D. and Steriade, M. (1991) Short-lasting nicotinic and long-lasting muscarinic depolarizing responses of thalamocortical neurons to stimulation of mesopontine cholinergic nuclei. Journal of Neurophysiology 65: 393–406.CrossRefGoogle ScholarPubMed
Curró Dossi, R., Nuñez, A. and Steriade, M. (1992a) Electrophysiology of a slow (0.5–4 Hz) intrinsic oscillation of cat thalamocortical neurones in vivo. Journal of Physiology (London) 447: 215–34.CrossRefGoogle Scholar
Curró Dossi, R., Paré, D. and Steriade, M. (1992b) Various types of inhibitory postsynaptic potentials in anterior thalamic cells are differentially altered by stimulation of laterodorsal tegmental cholinergic nucleus. Neuroscience 47: 279–89.CrossRefGoogle Scholar
Damasio, A. R. (1996) The somatic marker hypothesis and the possible functions of the prefrontal cortex. Philosophical Transactions of the Royal Society of LondonB351: 1413–20.CrossRefGoogle ScholarPubMed
Damasio, A. R. and Maurer, R. G. (1978) A neurological model for childhood autism. Archives of Neurology 35: 777–86.CrossRefGoogle ScholarPubMed
Danober, L. and Pape, H. C. (1998) Mechanisms and functional significance of a slow inhibitory potential in neurons of the lateral amygdala. European Journal of Neuroscience 10: 853–67.CrossRefGoogle ScholarPubMed
Datta, S., Curró Dossi, R., Paré, D., Oakson, G. and Steriade, M. (1991) Substantia nigra reticulata neurons during sleep-waking states: relation with ponto-geniculo-occipital waves. Brain Research 566: 344–7.CrossRefGoogle ScholarPubMed
Davenne, D. and Adrien, J. (1984) Suppression of PGO waves in the kitten: anatomical effects on the lateral geniculate nucleus. Neuroscience Letters 45: 33–8.CrossRefGoogle ScholarPubMed
Davis, M. (1992) The role of the amygdala in fear and anxiety. Annual Review of Neuroscience 15: 353–75.CrossRefGoogle ScholarPubMed
Davis, M. (2000) The role of the amygdala in conditioned and unconditioned fear and anxiety. In The Amygdala: A Functional Analysis, ed. Aggleton, J. P., pp. 213–87. Oxford: Oxford University Press.Google Scholar
Dawson, T., Bredt, D., Fotuhi, M., Hwang, P. and Snyder, S. (1991) Nitric oxide synthase and neuronal NADPH diaphorase are identical in brain and peripheral tissues. Proceedings of the National Academy of Sciences USA 88: 7797–801.CrossRefGoogle ScholarPubMed
Deacon, T. W., Eichenbaum, H., Rosenberg, P. and Eckman, K. W. (1983) Afferent connections of the perirhinal cortex in the rat. Journal of Comparative Neurology 220: 168–90.CrossRefGoogle ScholarPubMed
DeFelipe, J. and Fariñas, I. (1992) The pyramidal neuron of the cerebral cortex: morphological and chemical characteristics of the synaptic inputs. Progress in Neurobiology 39: 563–607.CrossRefGoogle ScholarPubMed
DeFelipe, J. and Jones, E. G. (1992) High-resolution light and electron microscopy immunocytochemistry of colocalized GABA and calbindin D-28k in somata and double bouquet cell axons in the monkey sensory-motor cortex. European Journal of Neuroscience 4: 46–60.CrossRefGoogle Scholar
DeFelipe, J., Elston, G. N., Fujita, I.et al. (2002) Neocortical circuits: evolutionary aspects and specificity versus non-specificity of synaptic connections. Remarks, main conclusions and general comments and discussion. Journal of Neurocytology 31: 387–416.CrossRefGoogle ScholarPubMed
Gennaro, L., Ferrara, M. and Bertini, M. (2001) The boundary between wakefulness and sleep: quantitative electroencephalographic changes during the sleep onset period. Neuroscience 107: 1–11.CrossRefGoogle ScholarPubMed
Demaurex, N., Lew, D. P. and Krause, K.-H. (1992) Cyclopiazonic acid depletes intracellular Ca2+ stores and activates an influx pathway for divalent cations in HL-60 cells. Journal of Biological Chemistry 267: 2318–24.Google ScholarPubMed
Dement, W. C., Ferguson, J., Cohen, H. and Barchas, J. (1969) Non-chemical methods and data using a biochemical model: the REM quanta. In Psychochemical Research in Man: Methods, Strategy and Theory, ed. Mandell, A. and Mandell, M. P., pp. 275–325. New York: Academic Press.Google Scholar
Denning, K. S. and Reinagel, P. (2005) Visual control of burst priming in the anesthetized lateral geniculate nucleus. Journal of Neuroscience 25: 3531–8.CrossRefGoogle ScholarPubMed
Oca, B. M., Cola, J. P., Maren, S. and Fanselow, M. S. (1998) Distinct regions of the periaqueductal gray are involved in the acquisition and expression if defensive responses. Journal of Neuroscience 18: 3426–32.CrossRefGoogle ScholarPubMed
Olmos, J. S. (1990) Amygdala. In The Human Nervous System, ed. Paxinos, G., pp. 583–708. New York: Academic Press.Google Scholar
Descarries, L., Watkins, K. C., and Lapierre, Y. (1977) Noradrenergic axon terminals in the cerebral cortex of rat. III. Topometric ultrastructural analysis. Brain Research 133: 197–222.CrossRefGoogle ScholarPubMed
Descarries, L., Gisiger, V. and Steriade, M. (1997) Diffuse transmission by acetylcholine in the CNS. Progress in Neurobiology 53: 603–25.CrossRefGoogle ScholarPubMed
Descarries, L., Mechawar, N., Aznavour, N. and Watkins, K. C. (2004) Structural determinants of the roles of acetylcholine in cerebral cortex. In Acetylcholine in the Cerebral Cortex (Progress in Brain Research, vol. 145), ed., Descarries, L., Krnjevic, K. and Steriade, M., pp. 45–58. Amsterdam: Elsevier.Google Scholar
Deschênes, M. (1981) Dendritic spikes induced in fast pyramidal tract neurons by thalamic stimulation. Experimental Brain Research 43: 304–8.Google ScholarPubMed
Deschênes, M. and Hu, B. (1990) Electrophysiology and pharmacology of the corticothalamic input to lateral thalamic nuclei: an intracellular study in the cat. European Journal of Neuroscience 2: 140–52.CrossRefGoogle Scholar
Deschênes, M. and Steriade, M. (1988) The neuronal mechanism of thalamic PGO waves. In Cellular Thalamic Mechanisms, ed. Bentivoglio, M., Macchi, G. and Spreafico, R., pp. 197–206. Amsterdam: Elsevier.Google Scholar
Deschênes, M., Roy, J. P. and Steriade, M. (1982) Thalamic bursting mechanism: an inward slow current revealed by membrane hyperpolarization. Brain Research 239: 289–93.CrossRefGoogle ScholarPubMed
Deschênes, M., Paradis, M., Roy, J. P. and Steriade, M. (1984) Electrophysiology of neurons of lateral thalamic nuclei in cat: resting properties and burst discharges. Journal of Neurophysiology 51: 1196–219.CrossRefGoogle ScholarPubMed
Deschênes, M., Madariaga-Domich, A. and Steriade, M. (1985) Dendrodendritic synapses in cat reticularis thalami nucleus, a structural basis for thalamic spindle synchronization. Brain Research 334: 169–71.Google ScholarPubMed
Deschênes, M., Veinante, P. and Zhang, Z. W. (1998).The organization of corticothalamic projections: reciprocity versus parity. Brain Research Reviews 28: 286–308.CrossRefGoogle ScholarPubMed
Desmedt, J. E. and Tomberg, C. (1994) Transient phase-locking of 40 Hz electrical oscillations in prefrontal and parietal human cortex reflects the process of conscious somatic perception. Neuroscience Letters 168: 126–9.CrossRefGoogle ScholarPubMed
Desmedt, J. E., Nguyen, T. H. and Bourget, M. (1983) The cognitive P40, N60 and P100 components of somatosensory evoked potentials and the earliest signs of sensory processing in man. Electroencephalography and Clinical Neurophysiology 56: 272–82.CrossRefGoogle ScholarPubMed
Destexhe, A. and Paré, D. (1999) Impact of network activity on the integrative properties of neocortical pyramidal neurons in vivo. Journal of Neurophysiology 81: 1531–47.CrossRefGoogle ScholarPubMed
Destexhe, A. and Paré, D. (2000) A combined computational and intracellular study of correlated synaptic bombardment in neocortical pyramidal neurons in vivo. Neurocomputing 32–3: 113–19.CrossRefGoogle Scholar
Destexhe, A., Contreras, D., Sejnowski, T. J. and Steriade, M. (1994a) A model of spindle rhythmicity in the isolated thalamic reticular nucleus. Journal of Neurophysiology 72: 803–18.CrossRefGoogle Scholar
Destexhe, A., Contreras, D., Sejnowski, T. J. and Steriade, M. (1994b) Modeling the control of reticular thalamic oscillations by neuromodulators. NeuroReport 5: 2217–20.CrossRefGoogle Scholar
Destexhe, A., Contreras, D., Steriade, M., Sejnowski, T. J. and Huguenard, J. R. (1996) In vivo, in vitro and computational analysis of dendritic calcium currents in thalamic reticular neurons. Journal of Neuroscience 16: 169–85.CrossRefGoogle ScholarPubMed
Destexhe, A., Contreras, D. and Steriade, M. (1999) Spatiotemporal analysis of local field potentials and unit discharges in cat cerebral cortex during natural wake and sleep states. Journal of Neuroscience 19: 4595–608.CrossRefGoogle ScholarPubMed
Destexhe, A., Contreras, D. and Steriade, M. (2001) LTS cells in cerebral cortex and their role in generating spike-and-wave oscillations. Neurocomputing 38–40: 555–63.CrossRefGoogle Scholar
Destexhe, A., Rudolph, M. and Paré, D. (2003) The high-conductance state of neocortical neurons in vivo. Nature Reviews Neuroscience 4: 739–51.CrossRefGoogle ScholarPubMed
Détári, L., Rasmusson, D. D. and Semba, K. (1997) Phasic relationship between the activity of basal forebrain neurons and cortical EEG in urethane-anesthetized rat. Brain Research 759: 112–21.CrossRefGoogle ScholarPubMed
Deuchars, J. and Thomson, A. M. (1995a) Single axon IPSPs elicited by a sparsely spiny interneuron in rat neocortex. Neuroscience 65: 935–42.CrossRefGoogle Scholar
Deuchars, J. and Thomson, A. M. (1995b) Innervation of burst firing spiny interneurons by pyramidal cells in deep layers of rat somatomotor cortex: paired intracellular recordings with biocytin filling. Neuroscience 69: 739–55.CrossRefGoogle Scholar
Dickinson, A. H., Mesches, M. H., Coleman, K. and McGaugh, J. L. (1993) Bicuculline administered into the amygdala blocks benzodiazepine-induced amnesia. Behavioral and Neural Biology 60: 1–4.CrossRefGoogle Scholar
Dickson, C. T., Biella, G. and Curtis, M. (2003) Slow periodic events and their transition to gamma oscillations in the entorhinal cortex of the isolated guinea pig brain. Journal of Neurophysiology 90: 39–46.CrossRefGoogle ScholarPubMed
Di Pasquale, E., Keegan, K. D. and Noebels, J. L. (1997) Increased excitability and inward rectification in layer V cortical pyramidal neurons in the epileptic mutant mouse Stargazer. Journal of Neurophysiology 77: 621–31.CrossRefGoogle Scholar
Dolan, R. J., Bench, C. J., Brown, R. G.et al. (1992) Regional cerebral blood flow abnormalities in depressed patients with cognitive impairment. Journal of Neurology, Neurosurgery and Psychiatry 55: 768–73.CrossRefGoogle ScholarPubMed
Dolmetsch, R. E., Pajvani, U., Fife, K., Spotts, J. M. and Greenberg, M. E. (2001) Signaling to the nucleus by an L-type calcium channel-calmodulin complex through the MAP kinase pathway. Science 294: 333–9.CrossRefGoogle ScholarPubMed
Domich, L., Oakson, G. and Steriade, M. (1986) Thalamic burst patterns in the naturally sleeping cat: a comparison between cortically projecting and reticularis neurones. Journal of Physiology (London) 379: 429–49.CrossRefGoogle ScholarPubMed
Douglas, R. J. and Martin, K. (1991) A functional microcircuit for cat visual cortex. Journal of Physiology (London) 440: 735–69.CrossRefGoogle ScholarPubMed
Douglas, R. J., Koch, C., Mahowald, M., Martin, K. A. C. and Suarez, H. H. (1995) Recurrent excitation in neocortical circuits. Science 269: 981–5.CrossRefGoogle ScholarPubMed
Doyère, V., Burette, F., Negro, C. R. and Laroche, S. (1993) Long-term potentiation of hippocampal afferents and efferents to prefrontal cortex: implications for associative learning. Neuropsychologia 31: 1031–53.CrossRefGoogle ScholarPubMed
Drevets, W. C. and Reichle, M. E. (1992) Neuroanatomical circuits in depression: implication for treatment mechanisms. Psychopharmacological Bulletin 28: 261–74.Google Scholar
Dringenberg, H. C. and Olmstead, M. C. (2003) Integrated contributions of basal forebrain and thalamus to neocortical activation elicited by pedunculopontine tegmental stimulation in urethane-anesthetized rats. Neuroscience 119: 839–53.CrossRefGoogle ScholarPubMed
Dringenberg, H. C. and Vanderwolf, C. H. (1996) Cholinergic activation of the electrocorticogram: an amygdaloid activating system. Experimental Brain Research 108: 285–96.CrossRefGoogle ScholarPubMed
Dumont, É.C., Martina, M., Samson, R. D., Drolet, G. and Paré, D. (2002) Physiological properties of central amygdala neurons: species differences. European Journal of Neuroscience 15: 545–52.CrossRefGoogle ScholarPubMed
Dumont, S. and Dell, P. (1960) Facilitation réticulaire des mécanismes visuels corticaux. Electroencephalography and Clinical Neurophysiology 12: 769–96.CrossRefGoogle Scholar
Duncan, C. P. (1949) The retroactive effect of electroshock on learning. Journal of Comparative Physiology and Psychology 42: 32–44.CrossRefGoogle ScholarPubMed
Duque, A., Balatoni, B., Détári, L. and Zaborsky, L. (2000) EEG correlation of the discharge properties of identified neurons in the basal forebrain. Journal of Neurophysiology 84: 1627–35.CrossRefGoogle ScholarPubMed
Durand, G. M., Kovalchuk, Y. and Konnerth, A. (1996) Long-term potentiation and functional synapse induction in developing hippocampus. Nature 381: 71–5.CrossRefGoogle ScholarPubMed
Eccles, J. C. (1961) Chairman's opening remarks. In The Nature of Sleep, ed. Wolstenholme, G. E. W. and O'Connor, M., pp. 1–3. London: Churchill.Google Scholar
Eckhorn, R., Bauer, R., Jordan, W.et al. (1988) Coherent oscillations: a mechanism of feature linking in the visual cortex?Biological Cybernetics 60: 121–30.CrossRefGoogle ScholarPubMed
Egan, T. M. and North, R. A. (1985) Acetylcholine acts on m2-muscarinic receptors to excite rat locus coeruleus neurones. British Journal of Pharmacology 85: 733–5.CrossRefGoogle ScholarPubMed
Ehlers, C., Hendricksen, S. J., Wang, M.et al. (1983) Corticotropin releasing factor produces increases in brain excitability and convulsive seizures in rats. Brain Research 278: 332–6.CrossRefGoogle ScholarPubMed
Eichenbaum, H. (2002) Memory representations in the parahippocampal region. In The Parahippocampal Region, ed. Witter, M. P. and Wouterlood, F., pp. 165–84. Oxford: Oxford University Press.Google Scholar
Eichenbaum, H., Schoenbaum, G., Young, B. and Bunsey, M. (1996) Functional organization of the hippocampal memory system. Proceedings of the National Academy of Sciences USA 93: 13500–7.CrossRefGoogle ScholarPubMed
El Mansari, M., Sakai, K. and Jouvet, M. (1989) Unitary characteristics of presumptive cholinergic tegmental neurons during the sleep-waking cycle in freely moving cats. Experimental Brain Research 76: 519–29.CrossRefGoogle ScholarPubMed
Emerson, R. G., Sgro, J. A., Pedley, T. A. and Hauser, A. (1988) State-dependent changes in the N20 component of the median nerve somatosensory evoked potential. Neurology 38: 64–7.CrossRefGoogle ScholarPubMed
Emptage, N., Bliss, T. and Fine, A. (1999) Single synaptic events evoke NMDA receptor-mediated release of calcium from internal stores in hippocampal dendritic spines. Neuron 22: 115–24.CrossRefGoogle ScholarPubMed
Endo, K., Araki, T. and Ito, K. (1977) Short latency EPSPs and incrementing PSPs of pyramidal tract cells evoked by stimulation of the nucleus centralis lateralis of the thalamus. Brain Research 132: 541–6.CrossRefGoogle ScholarPubMed
Engel, A. K., König, P., Kreiter, A. K. and Singer, W. (1991).Interhemispheric synchronization of oscillatory neuronal responses in cat visual cortex. Science 252: 1177–9.CrossRefGoogle ScholarPubMed
Engel, A. K., Fries, P and Singer, W. (2001) Dynamic predictions: oscillations and synchrony in top-down processing. Nature Reviews Neuroscience 2: 704–16.CrossRefGoogle ScholarPubMed
Esclapez, M., Tillakaratne, N., Tobin, A. J. and Houser, C. R. (1993) Comparative localization of two forms of glutamic acid decarboxylase with nonradioactive in situ hybridization methods. Journal of Comparative Neurology 331: 339–62.CrossRefGoogle ScholarPubMed
Evarts, E. V. (1964) Temporal patterns of discharge of pyramidal tract neurons during sleep and waking. Journal of Neurophysiology 27: 152–71.CrossRefGoogle ScholarPubMed
Evarts, E. V. (1965) Relation of discharge frequency to conduction velocity in pyramidal tract neurons. Journal of Neurophysiology 28: 216–28.CrossRefGoogle ScholarPubMed
Faber, E. and Sah, P. (2002) Physiological role of calcium-activated potassium currents in the rat lateral amygdala. Journal of Neuroscience 22: 1618–28.CrossRefGoogle ScholarPubMed
Faber, E. S. and Sah, P. (2003) Calcium-activated K+ (BK) channel inactivation contributes to spike broadening during repetitive firing in rat lateral amygdala neurons. Journal of Physiology (London) 552 (2): 482–97.CrossRefGoogle Scholar
Faber, E., Callister, R. J. and Sah, P. (2001) Morphological and electrophysiological properties of principal neurons in the rat lateral amygdala in vitro. Journal of Neurophysiology 85: 714–23.CrossRefGoogle ScholarPubMed
Façon, E., Steriade, M. and Wertheimer, N. (1958) Hypersomnie prolongée engendrée par des lésions bilatérales du système activateur médial: le syndrome thrombotique de la bifurcation du tronc basilaire. Revue Neurologique (Paris) 98: 117–33.Google Scholar
Fahy, F. L., Riches, I. P. and Brown, M. W. (1993) Neuronal activity related to visual recognition memory: long-term memory and the encoding of recency and familiarity information in the primate anterior and medial inferior temporal cortex and rhinal cortex. Experimental Brain Research 96: 457–72.CrossRefGoogle ScholarPubMed
Fallon, J. H. and Ciofi, P. (1992) Distribution of monoamines with the amygdala. In The Amygdala, ed. Aggleton, J. P., pp. 97–114. New York: Wiley-Liss.Google ScholarPubMed
Fallon, J. H., Koziell, D. A. and Moore, R. Y. (1978) Catecholamine innervation of the basal forebrain. II. Amygdala, suprarhinal cortex and entorhinal cortex. Journal of Comparative Neurology 180: 509–32.CrossRefGoogle ScholarPubMed
Falls, W. A., Miserendino, M. J. D. and Davis, M. (1992) Extinction of fear-potentiated startle: blockade by infusion of an NMDA antagonist into the amygdala. Journal of Neuroscience 12: 854–63.CrossRefGoogle ScholarPubMed
Fanselow, M. S. and Kim, J. J. (1994) Acquisition of contextual Pavlovian fear conditioning is blocked by application of an NMDA receptor antagonist D,L-2-amino-5-phosphonovaleric acid to the basolateral amygdala. Behavioral Neuroscience 108: 210–12.CrossRefGoogle ScholarPubMed
Fanselow, M. S. and LeDoux, J. E. (1999) Why we think plasticity underlying Pavlovian fear conditioning occurs in the basolateral amygdala. Neuron 23: 229–32.CrossRefGoogle ScholarPubMed
Fanselow, M. S., Kim, J. J., Yipp, J. and Oca, B. (1994) Differential effects of the N-methyl-D-aspartate antagonist DL-2-amino-5-phosphonovalerate on acquisition of fear of auditory and contextual cues. Behavioral Neuroscience 108: 235–40.CrossRefGoogle ScholarPubMed
Farb, C. R. and LeDoux, J. E. (1999) Afferents from rat temporal cortex synapse on lateral amygdala neurons that express NMDA and AMPA receptors. Synapse 33: 218–29.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Farmer, S. F. (1998) Rhythmicity, synchronization and binding in human primate cortex. Journal of Physiology (London) 509: 3–14.CrossRefGoogle Scholar
Faulkner, B. and Brown, T. H. (1999) Morphology and physiology of neurons in the rat perirhinal-lateral amygdala area. Journal of Comparative Neurology 411: 613–42.3.0.CO;2-U>CrossRefGoogle ScholarPubMed
Feinberg, I. and Campbell, I. G. (1993) Ketamine administration during waking increases delta EEG intensity in rat sleep. Neuropharmacology 9: 41–8.Google ScholarPubMed
Fellous, J. M., Rudolph, M., Destexhe, A. and Sejnowski, T. J. (2003) Synaptic background noise controls the input/output characteristics of single cells in an in vitro model of in vivo activity. Neuroscience 122: 811–29.CrossRefGoogle Scholar
Ferrara, M., Gennaro, L. and Bertini, M. (1999) Selective slow-wave sleep (SWS) deprivation and SWS rebound: do we need a fixed amount per night?Sleep Research Online 2: 15–19.Google Scholar
Ferrara, M., Gennaro, L., Curcio, G.et al. (2002) Regional differences of the human sleep electroencephalogram in response to selective slow-wave sleep deprivation. Cerebral Cortex 12: 737–48.CrossRefGoogle ScholarPubMed
Ferry, B. and McGaugh, J. L. (1999) Clenbuterol administration into the basolateral amygdala post-training enhances retention in an inhibitory avoidance task. Neurobiology of Learning and Memory 72: 8–12.CrossRefGoogle Scholar
Ferry, B., Magistretti, P. J. and Pralong, E. (1997) Noradrenaline modulates glutamate-mediated neurotransmission in the rat basolateral amygdala in vitro. European Journal of Neuroscience 9: 1356–64.CrossRefGoogle ScholarPubMed
Feshchenko, V. A. and Chilingaryan, L. I. (1990) Dependence of electrical activity of the amygdaloid complex on level of motivation and emotional state of the dog. Neuroscience Behavioral Physiology 20: 506–13.CrossRefGoogle ScholarPubMed
Finch, D. M., Wong, E. E., Derian, E. L., Chen, X. H., Nowlin-Finch, N. L. and Brothers, L. A. (1986) Neurophysiology of limbic system pathways in the rat: projections from the amygdala to the entorhinal cortex. Brain Research 370: 273–84.CrossRefGoogle ScholarPubMed
Finch, D. M., Tan, A. M. and Isokawa-Akesson, M. (1988) Feedforward inhibition of the rat entorhinal cortex and subicular complex. Journal of Neuroscience 8: 2213–26.CrossRefGoogle ScholarPubMed
Finelli, L. A., Baumann, H., Borbély, A. and Achermann, P. (2000) Dual electroencephalogram markers of human sleep homeostasis: correlation between theta activity in waking and slow-wave activity in sleep. Neuroscience 101: 523–9.CrossRefGoogle ScholarPubMed
Fiset, P., Paus, T., Daloze, T.et al. (1999) Brain mechanisms of propofol-induced loss of consciousness in humans: a positron emission tomography study. Journal of Neuroscience 19: 5506–13.CrossRefGoogle Scholar
Foehring, R. C. and Waters, R. S. (1991) Contributions of low-threshold calcium current and anomalous rectifier (Ih) to slow depolarizations underlying burst firing in human neocortical neurons in vitro. Neuroscience Letters 124: 17–21.CrossRefGoogle ScholarPubMed
Foehring, R. C., Schwindt, P. C. and Crill, W. E. (1989) Norepinephrine selectively reduces slow Ca2+- and Na+ -mediated K+ currents in cat neocortical neurons. Journal of Neurophysiology 61: 245–256.CrossRefGoogle ScholarPubMed
Fonnum, S., Strom-Mathisen, J. and Divac, I. (1981) Biochemical evidence for glutamate as neurotransmitter in corticostriatal and corticothalamic fibres in rat brain. Neuroscience 6: 863–73.CrossRefGoogle ScholarPubMed
Fontanini, A., Spano, P and Bower, J. M. (2003) Ketamine-xylazine-induced slow (<1.5 Hz) oscillations in the rat piriform (olfactory) cortex are functionally correlated with respiration. Journal of Neuroscience 23: 7993–8001.CrossRefGoogle ScholarPubMed
Foote, S. L., Bloom, F. E., and Aston-Jones, G. (1983) Nucleus locus ceruleus: new evidence of anatomical and physiological specificity. Physiological Reviews 63: 844–914.CrossRefGoogle ScholarPubMed
Ford, B., Holmes, C. J., Mainville, L. and Jones, B. E. (1995) GABAergic neurons in the rat pontomesencephalic tegmentum: codistribution with cholinergic and other tegmental neurons projecting to the posterior lateral hypothalamus. Journal of Comparative Neurology 363: 177–96.CrossRefGoogle ScholarPubMed
Forslid, A., Andersson, B. and Johansson, S. (1986) Observations on normal EEG activity in different brain regions of the unrestrained swine. Acta Physiologica Scandinavica 128: 389–96.CrossRefGoogle ScholarPubMed
Fort, P., Khateb, A., Pegna, A., Jones, B. E. and Mühlethaler, M. (1995) Noradrenergic modulation of cholinergic nucleus basalis neurons demonstrated by in vitro pharmacological and immunohistochemical evidence in the guinea pig brain. European Journal of Neuroscience 7: 1502–11.CrossRefGoogle ScholarPubMed
Fortin, M., Asselin, M. C., Gould, P. V. and Parent, A. (1998) Calretinin-immunoreactive neurons in the human thalamus. Neuroscience 84: 537–48.CrossRefGoogle ScholarPubMed
Fosse, R., Stickgold, R. and Hobson, J. A. (2004) Thinking and hallucinating: reciprocal changes in sleep. Psychophysiology 41: 298–305.CrossRefGoogle ScholarPubMed
Foster, J. A. (1980) Intracortical origin of recruiting responses in the cat neocortex. Electroencephalography and Clinical Neurophysiology 48: 639–53.CrossRefGoogle Scholar
Foulkes, D. (1962) Dream reports from different stages of sleep. Journal of Abnormal and Social Psychology 65: 14–25.CrossRefGoogle Scholar
Foulkes, D. and Schmidt, M. (1983) Temporal sequence and unit composition in dream reports from different stages of sleep. Sleep 6: 265–80.CrossRefGoogle Scholar
Fournier, G. N., Materi, L. M., Semba, K. and Rasmusson, D. D. (2004) Cortical acetylcholine release and electroencephalogram activation evoked by ionotropic glutamate receptor agonists in the rat basal forebrain. Neuroscience 123: 785–92.CrossRefGoogle ScholarPubMed
Fox, C. A. (1940) Certain basal telencephalic centers in the cat. Journal of Comparative Neurology 72: 1–62.CrossRefGoogle Scholar
Frank, M. G., Issa, N. P. and Stryker, M. P. (2001) Sleep enhances plasticity in the developing visual cortex. Neuron 30: 275–87.CrossRefGoogle ScholarPubMed
Freedman, J. E. and Aghajanian, G. K. (1987) Role of phosphoinositide metabolites in the prolongation of afterhyperpolarizations by alpha 1-adrenoceptors in rat dorsal raphe neurons. Journal of Neuroscience 7: 3897–906.CrossRefGoogle ScholarPubMed
Freedman, L. J., Insel, T. R. and Smith, Y. (2000) Subcortical projections of area 25 (subgenual cortex) of the macaque monkey. Journal of Comparative Neurology 29: 172–88.3.0.CO;2-8>CrossRefGoogle Scholar
Freeman, W. J. (1959) Distribution in time and space of prepyriform electrical activity. Journal of Neurophysiology 22: 644–65.CrossRefGoogle ScholarPubMed
Freeman, W. J. (1975) Mass Action in the Nervous System. New York: Academic Press.Google Scholar
Freeman, W. J. and Va, Dijk, B. W. (1988) Spatial patterns of visual cortical fast EEG during conditioned reflex in a rhesus monkey. Brain Research 422: 267–76.CrossRefGoogle Scholar
Frégnac, Y. (1999) A tale of two spikes. Nature Neuroscience 2: 299–301.CrossRefGoogle ScholarPubMed
Freund, T. F. and Buzsáki, G. (1996) Interneurons of the hippocampus. Hippocampus 6: 347–470.3.0.CO;2-I>CrossRefGoogle ScholarPubMed
Freund, T. F. and Meskenaite, V. (1992) γ-Aminobutyric acid-containing basal forebrain neurons innervate inhibitory interneurons in the neocortex. Proceedings of the National Academy of Sciences USA 89: 738–42.CrossRefGoogle ScholarPubMed
Friedman, A. and Gutnick, M. J. (1989) Intracellular calcium and control of burst generation in neurons of guinea-pig neocortex in vitro. European Journal of Neuroscience 1: 374–81.CrossRefGoogle ScholarPubMed
Fries, P., Reynolds, J. H., Rorie, A. E. and Desimone, R. (2001) Modulation of oscillatory neuronal synchronization by selective visual attention. Science 291: 1560–3.CrossRefGoogle ScholarPubMed
Fu, Y. and Shinnick-Gallagher, P. (2005) Two intra-amygdaloid pathways to the central amygdala exhibit different mechanisms of long-term potentiation. Journal of Neurophysiology 93: 3012–15.CrossRefGoogle ScholarPubMed
Fuentealba, P. and Steriade, M. (2005) The reticular nucleus revisited: intrinsic and network properties of a thalamic pacemaker. Progress in Neurobiology 75: 125–41.CrossRefGoogle ScholarPubMed
Fuentealba, P., Crochet, S., Timofeev, I. and Steriade, M. (2002) ‘Spikelets’ in cat thalamic reticular nucleus in vivo. Society for Neuroscience Abstracts 28: 144.19.Google Scholar
Fuentealba, P., Crochet, S. and Steriade, M. (2004a) The cortically evoked secondary depolarization affects the integrative properties of thalamic reticular neurons. European Journal of Neuroscience 20: 2691–6.CrossRefGoogle Scholar
Fuentealba, P., Crochet, S., Timofeev, I.et al. (2004b) Experimental evidence and modeling studies support a synchronizing role for electrical coupling in the cat thalamic reticular neurons in vivo. European Journal of Neuroscience 20: 111–19.CrossRefGoogle Scholar
Fuentealba, P., Crochet, S., Timofeev, I. and Steriade, M. (2004c) Synaptic interactions between thalamic and cortical inputs onto cortical neurons in vivo. Journal of Neurophysiology 91: 1990–8.CrossRefGoogle Scholar
Fuentealba, P., Timofeev, I. and Steriade, M. (2004d) Prolonged hyperpolarizing potentials precede spindle oscillations in the thalamic reticular nucleus. Proceedings of the National Academy of Sciences USA 101: 9816–21.CrossRefGoogle Scholar
Fuentealba, P., Timofeev, I., Bazhenov, M., Sejnowski, T. J. and Steriade, M. (2005) Membrane bistability in cat thalamic reticular neurons during spindle oscillations. Journal of Neurophysiology 93: 294–304.CrossRefGoogle Scholar
Fuster, J. M. (2004) Upper processing stages of the perception-action cycle. Trends in Cognitive Sciences 8: 143–5.CrossRefGoogle ScholarPubMed
Fuxe, K., Jacobsen, K. X., Hoistad, M.et al. (2003) The dopamine D1 receptor-rich main and paracapsular intercalated nerve cell groups of the rat amygdala: relationship to the dopamine innervation. Neuroscience 119: 733–46.CrossRefGoogle ScholarPubMed
Gaffan, D. and Murray, E. A. (1992) Monkeys (Macaca fascicularis) with rhinal cortex ablations succeed in object discrimination learning despite 24-hr intertrial intervals and fail at matching to sample despite double sample presentations. Behavioral Neuroscience 106: 30–8.CrossRefGoogle ScholarPubMed
Gais, S. and Born, J. (2004) Low acetylcholine during slow-wave sleep is critical for declarative memory consolidation. Proceedings of the National Academy of Sciences USA 101: 2140–4.CrossRefGoogle ScholarPubMed
Gais, S., Plihal, W., Wagner, U. and Born, J. (2000) Early sleep triggers memory for early visual discrimination skills. Nature Neuroscience 3: 1335–9.CrossRefGoogle ScholarPubMed
Gais, S., Mölle, M., Helms, K. and Born, J. (2002) Learning-dependent increases in sleep density. Journal of Neuroscience 22: 6830–4.CrossRefGoogle Scholar
Galarreta, M. and Hestrin, S. (1999) A network of fast-spiking cells in the neocortex connected by electrical synapses. Nature 402: 72–5.CrossRefGoogle ScholarPubMed
Galarreta, M. and Hestrin, S. (2002) Electrical and chemical synapses among parvalbumin fast-spiking GABAergic interneurons in adult mouse neocortex. Proceedings of the National Academy of Sciences USA 99: 12438–43.CrossRefGoogle ScholarPubMed
Gallagher, M., Kapp, B. S., Musty, R. E. and Driscoll, P. A. (1977) Memory formation: evidence for a specific neurochemical system in the amygdala. Science 198: 423–5.CrossRefGoogle ScholarPubMed
Garcia-Rill, E., Biedermann, J. A., Chambers, T.et al. (1995) Mesopontine neurons in schizophrenia. Neuroscience 66: 321–35.CrossRefGoogle Scholar
Gaudreau, H. and Paré, D. (1996) Projection cells of the lateral nucleus are virtually silent throughout the sleep-waking cycle. Journal of Neurophysiology 75: 1301–5.CrossRefGoogle ScholarPubMed
Gaykema, R. P., Luiten, P. G., Nyakas, C. and Traber, J. (1990) Cortical projection patterns of the medial septum-diagonal band complex. Journal of Comparative Neurology 293: 103–24.CrossRefGoogle ScholarPubMed
Gean, P. W., Huang, C. C., Lin, J. H. and Tsai, J. J. (1992) Sustained enhancement of NMDA receptor-mediated synaptic potential by isoproterenol in rat amygdalar slices. Brain Research 594: 331–4.CrossRefGoogle ScholarPubMed
Gemmell, C. and O'Mara, S. M. (2000) Long-term potentiation and paired-pulse facilitation in the prelimbic cortex of the rat following stimulation in the contralateral hemisphere in vivo. Experimental Brain Research 132: 223–9.CrossRefGoogle ScholarPubMed
Gerard, R. W. (1949) Physiology and psychiatry. American Journal of Psychiatry 106: 161–73.CrossRefGoogle ScholarPubMed
German, D. C., Manaye, K. F., Wu, D., Hersh, L. B. and Zweig, R. M. (1999) Mesopontine cholinergic and non-cholinergic neurons in schizophrenia. Neuroscience 94: 33–8.CrossRefGoogle Scholar
Gervasoni, D., Lin, S. C., Ribeiro, S.et al. (2004) Global forebrain dynamics predict rat behavioral states and their transitions. Journal of Neuroscience 24: 11137–47.CrossRefGoogle ScholarPubMed
Gewirtz, J. C., Falls, W. A. and Davis, M. (1997) Normal conditioned inhibition and extinction of freezing and fear-potentiated startle following electrolytic lesions of medical prefrontal cortex in rats. Behavioral Neuroscience 111: 712–26.CrossRefGoogle ScholarPubMed
Ghashghaei, H. T. and Barbas, H. (2002) Pathways for emotion: Interactions of prefrontal and anterior. Neuroscience 115: 1261–79.CrossRefGoogle ScholarPubMed
Gibson, J. R., Beierlein, M. and Connors, B. W. (1999) Two networks of electrically coupled inhibitory neurons in neocortex. Nature 402: 75–9.CrossRefGoogle ScholarPubMed
Gilbert, C. D. (1998) Adult cortical dynamics. Physiological Reviews 78: 467–85.CrossRefGoogle ScholarPubMed
Girod, R., Barazangi, N., McGehee, D. and Role, L. W. (2000) Facilitation of glutamatergic neurotransmission by presynaptic nicotinic acetylcholine receptors. Neuropharmacology 39: 2715–25.CrossRefGoogle ScholarPubMed
Glenn, L. L. and Steriade, M. (1982) Discharge rate and excitability of cortically projecting intralaminar thalamic neurons during waking and sleep states. Journal of Neuroscience 2: 1287–404.CrossRefGoogle ScholarPubMed
Glenn, L. L., Hada, J., Roy, J. P., Deschênes, M. and Steriade, M. (1982) Anterograde tracer and field potential analysis of the neocortical layer I projection from nucleus ventralis medialis of the thalamus in cat. Neuroscience 7: 1861–77.CrossRefGoogle ScholarPubMed
Gloor, P. (1976) Generalized and widespread bilateral paroxysmal abnormalities. In Handbook of Electroencephalography and Clinical Neurophysiology, vol. 11, Part B, ed. Remond, A., pp. 52–87. Amsterdam: Elsevier.Google Scholar
Gold, P. E. and McGaugh, J. L. (1975) A single-trace, two process view of memory storage processes. In Short-term Memory, ed. Deutsch, D. and Deutsch, J. A., pp. 355–78. New York: Academic Press.Google Scholar
Goldman, P. S. (1979) Contralateral projections to the dorsal thalamus from frontal association cortex in rhesus monkey. Brain Research 166: 166–71.CrossRefGoogle ScholarPubMed
Goldman-Rakic, P. S. (1987) Circuitry of the prefrontal cortex and the regulation of behavior by representational memory. In Handbook of Physiology, vol. 5 (The Nervous System), ed. Plum, F. and Mountcastle, V. B., pp. 373–417. Bethesda, MD: American Physiological Society.Google Scholar
Goldman-Rakic, P. S. (1988) Changing concepts of cortical connectivity: parallel distributed cortical networks. In Neurobiology of Neocortex, ed. Kakic, P. and Singer, W., pp. 177–202. New York: Wiley.Google Scholar
Golomb, D., Wang, X. J. and Rinzel, J. (1994) Synchronization properties of spindle oscillations in a thalamic reticular nucleus model. Journal of Neurophysiology 72: 1109–26.CrossRefGoogle Scholar
Golshani, P., Liu, X. B. and Jones, E. G. (2001) Differences in quantal amplitude reflect GluR4-subunit number at corticothalamic synapses on two populations of thalamic neurons. Proceedings of the National Academy of Sciences USA 98: 4172–7.CrossRefGoogle ScholarPubMed
Goosens, G. and Otto, T. A. (1998) Induction and transient suppression of long-term potentiation in the peri- and postrhinal cortices following theta-related stimulation of hippocampal field CA1. Brain Research 780: 95–101.Google Scholar
Goosens, K. A. and Maren, S. (2001) Contextual and auditory fear conditioning are mediated by the lateral, basal, and central amygdaloid nuclei in rats. Learning and Memory 8: 148–55.CrossRefGoogle ScholarPubMed
Goosens, K. A., and Maren, S. (2003) Pretraining NMDA receptor blockade in the basolateral complex. Behavioral Neuroscience 117: 738–50.CrossRefGoogle ScholarPubMed
Govindaiah and Cox, C. L. (2004) Synaptic activation of metabotropic glutamate receptors regulates dendritic outputs of thalamic interneurons. Neuron 41: 611–23.CrossRefGoogle Scholar
Grace, A. A. and Bunney, B. S. (1979) Paradoxical GABA excitation of nigral dopaminergic cells: indirect mediation through reticulata inhibitory neurons. European Journal of Pharmacology 59: 211–18.CrossRefGoogle ScholarPubMed
Grace, A. A. and Bunney, B. S. (1985) Opposing effects of striatonigral feed-back pathways on midbrain dopamine cell activity. Brain Research 333: 271–84.CrossRefGoogle Scholar
Grastyan, E., Lissak, K., Madarasz, I. and Donhoffer, H. (1959) The hippocampal electrical activity during the development of conditioned reflexes. Electroencephalography and Clinical Neurophysiology 11: 409–30.CrossRefGoogle ScholarPubMed
Gray, C. M. and McCormick, D. A. (1996) Chattering cells: superficial pyramidal neurons contributing to the generation of synchronous oscillations in the visual cortex. Science 274: 109–13.CrossRefGoogle ScholarPubMed
Gray, C. M., König, P., Engel, A. K. and Singer, W. (1989) Stimulus-specific neuronal oscillations in cat visual cortex exhibit inter-columnar synchronization which reflects global stimulus properties. Nature 338: 334–7.CrossRefGoogle Scholar
Gray, C. M., Engel, A. K., König, P. and Singer, W. (1990) Stimulus-dependent neuronal oscillations in cat visual cortex: receptive field properties and feature dependence. European Journal of Neuroscience 2: 607–19.CrossRefGoogle ScholarPubMed
Graybiel, A. M. and Ragsdale, C. W. (1979) Fiber connections of the basal ganglia. In Development and Chemical Specificity of Neurons, ed. Cuénod, M., Kreutzberg, G. W., and Bloom, F. R., pp. 239–83. Amsterdam: Elsevier.
Green, J. D. and Arduini, A. (1954) Hippocampal electrical activity in arousal. Journal of Neurophysiology 17: 533–57.CrossRefGoogle ScholarPubMed
Grenier, F., Timofeev, I. and Steriade, M. (2001) Focal synchronization of ripples (80–200 Hz) in neocortex and their neuronal correlates. Journal of Neurophysiology 86: 1884–98.CrossRefGoogle ScholarPubMed
Griffith, W. H. (1988) Membrane properties of cell types within guinea pig basal forebrain nuclei in vitro. Journal of Neurophysiology 59: 1590–612.CrossRefGoogle ScholarPubMed
Gritti, I., Mainville, L. and Jones, B. E. (1994) Projections of GABAergic and cholinergic basal forebrain and GABAergic preoptic-anterior hypothalamic neurons to the posterior lateral hypothalamus of the rat. Journal of Comparative Neurology 339: 251–68.CrossRefGoogle ScholarPubMed
Gross, D. W. and Gotman, J. (1999) Correlation of high-frequency oscillations with the sleep-wake cycle and cognitive activity in humans. Neuroscience 94: 1005–18.CrossRefGoogle ScholarPubMed
Guido, W. and Weyand, T. (1995) Burst responses in thalamic relay cells of the awake, behaving cat. Journal of Neurophysiology 74: 1782–6.CrossRefGoogle ScholarPubMed
Guido, W., Lu, S. M. and Sherman, S. M. (1992) Relative contributions of burst and tonic responses to the receptive field properties of lateral geniculate neurons in the cat. Journal of Neurophysiology 68: 2199–211.CrossRefGoogle ScholarPubMed
Guillery, R. W. (1969) The organization of synaptic interconnections in the laminae of the dorsal lateral geniculate nucleus of the cat. Zeitschrift für Zellforschung und Mikroskopische Anatomie 96: 1–38.CrossRefGoogle ScholarPubMed
Guillery, R. W. and Harting, J. K. (2002) Structure and connections of the thalamic reticular nucleus: advancing views over half of century. Journal of Comparative Neurology 463: 360–71.CrossRefGoogle Scholar
Gulyás, A. I., Hájos, N. and Freund, T. F. (1996) Interneurons containing calretinin are specialized to control other interneurons in the rat hippocampus. Journal of Neuroscience 16: 3397–411.CrossRefGoogle ScholarPubMed
Gupta, A., Wang, Y. and Markram, H. (2000) Organizing principles for a diversity of GABAergic interneurons and synapses in the neocortex. Science 287: 273–8.CrossRefGoogle ScholarPubMed
Gurden, H., Tassin, J. P. and Jay, T. M. (1999) Integrity of the mesocortical dopaminergic system is necessary for complete expression of in vivo hippocampal-prefrontal cortex long-term potentiation. Neuroscience 94: 1019–27.CrossRefGoogle ScholarPubMed
Gutnick, M. J., Heinemann, U. and Prince, D. A. (1979) Stimulus induced and seizure related changes in extracellular potassium concentration in cat thalamus (VPL). Electroencephalography and Clinical Neurophysiology 47: 329–44.CrossRefGoogle Scholar
Halgren, E., Bab, T. L. and Crandall, P. H. (1978) Human hippocampal formation EEG desynchronizes during attentiveness and movement. Electroencephalography and Clinical Neurophysiology 44: 778–81.CrossRefGoogle ScholarPubMed
Halgren, E., Boujon, C., Clarke, J., Wang, C. and Chauvel, P. (2002) Rapid distributed fronto-parieto-occipital processing stages during working memory in humans. Cerebral Cortex 12: 710–28.CrossRefGoogle ScholarPubMed
Hall, E. (1972a) The amygdala of the cat: A Golgi study. Zeitschrift für Zellforschung 134: 439–58.CrossRefGoogle Scholar
Hall, E. (1972b) Some aspects of the structural organization of the amygdala. In The Neurobiology of the Amygdala, ed. Eleftheriou, B. E., pp. 95–121. New York: Plenum Press.CrossRefGoogle Scholar
Hall, G. B. C., Szechtman, H. and Nahmias, C. (2003) Enhanced salience and emotion recognition in autism: a PET study. American Journal of Psychiatry 160: 1439–41.CrossRefGoogle ScholarPubMed
Halliwell, J. V. (1986) M-current in human neocortical neurones. Neuroscience Letters 67: 1–6.CrossRefGoogle ScholarPubMed
Hamam, B. N., Kennedy, T. E., Alonso, A. and Amaral, D. G. (2000) Morphological and electrophysiological characteristics of layer V neurons of the rat medial entorhinal cortex. Journal of Comparative Neurology 418: 457–72.3.0.CO;2-L>CrossRefGoogle Scholar
Hamann, S. B., Ely, T. D., Grafton, S. T. and Kilts, C. D. (1999) Amygdala activity related to enhanced memory for pleasant and aversive stimuli. Nature Neuroscience 2: 289–93.CrossRefGoogle ScholarPubMed
Hammond, C. and Crépel, F. (1992) Evidence for a slowly inactivating K+ current in prefrontal cortical cells. European Journal of Neuroscience 4: 1087–92.CrossRefGoogle ScholarPubMed
Hardy, S. G. and Leichnetz, G. R. (1981) Frontal cortical projections to the periaqueductal gray in the rat: a retrograde and orthograde horseradish peroxidase study. Neuroscience Letters 23: 13–17.CrossRefGoogle ScholarPubMed
Hartveit, E. and Heggelund, P. (1994) Response variability of single cells in the dorsal lateral geniculate nucleus of the cat. Comparison with retinal input and effect of brain stem stimulation. Journal of Neurophysiology 72: 1278–89.CrossRefGoogle ScholarPubMed
Hashikawa, T., Rauseell, E., Molinari, M. and Jones, E. G. (1991) Parvalbumin- and calbindin-containing neurons in the monkey medial geniculate complex: differential distribution and cortical layer specific projections. Brain Research 544: 335–41.CrossRefGoogle ScholarPubMed
Hasselmo, M. E. (1999) Neuromodulation: acetylcholine and memory consolidation. Trends in Cognitive Sciences 3: 351–9.CrossRefGoogle ScholarPubMed
Hasue, R. H. and Shammah-Lagnado, S. J. (2002) Origin of the dopaminergic innervation of the central extended amygdala and accumbens shell: a combined retrograde tracing and immunohistochemical study in the rat. Journal of Comparative Neurology 454: 15–33.CrossRefGoogle ScholarPubMed
Hatfield, T. and McGaugh, J. L. (1999) Norepinephrine infused into the basolateral amygdala posttraining enhances retention in a spatial water maze task. Neurobiology of Learning and Memory 71: 232–9.CrossRefGoogle Scholar
He, J. (2003) Slow oscillation in non-lemniscal auditory thalamus. Journal of Neuroscience 23: 8281–90.CrossRefGoogle ScholarPubMed
Herbert, M. R., Ziegler, D. A., Deutsch, C. K.et al. (2003) Dissociations of cerebral cortex, subcortical and cerebral white matter volumes in autistic boys. Brain 126: 1182–92.CrossRefGoogle ScholarPubMed
Herculano-Houzel, S., Munk, M. H. J., Neuenschwander, S. and Singer, W. (1999) Precisely synchronized oscillatory firing patterns require electroencephalographic activation. Journal of Neuroscience 19: 3992–4010.CrossRefGoogle ScholarPubMed
Herkenham, M. (1978) The connections of the nucleus reuniens thalami: evidence for a direct thalamo-hippocampal pathway in the rat. Journal of Comparative Neurology 177: 589–610.CrossRefGoogle ScholarPubMed
Herkenham, M. (1979) The afferent and efferent connections of the ventromedial thalamic nucleus in the rat. Journal of Comparative Neurology 183: 487–518.CrossRefGoogle ScholarPubMed
Hernández-Cruz, A. and Pape, H. C. (1989) Identification of two calcium currents in acutely dissociated neurons from the rat lateral geniculate nucleus. Journal of Neurophysiology 61: 1270–83.CrossRefGoogle ScholarPubMed
Herry, C. and Garcia, R. (2003) Behavioral and paired-pulse facilitation analyses of long-lasting depression at excitatory synapses in the medial prefrontal cortex in mice. Behavioral Brain Research 146: 89–96.CrossRefGoogle ScholarPubMed
Hestrin, S. and Armstrong, W. E. (1996) Morphology and physiology of cortical neurons in layer I. Journal of Neuroscience 16: 5290–300.CrossRefGoogle ScholarPubMed
Hicks, T. P., Metherate, R., Landry, P. and Dykes, R. W. (1986) Bicuculline-induced alterations of response properties in functionally identified ventroposteiror thalamic neurones. Experimental Brain Research 63: 248–64.CrossRefGoogle ScholarPubMed
Hikosaka, O. and Wurtz, R. H. (1983) Visual and oculomotor function of monkey substantia nigra pars reticulata, IV: relation of substantia nigra to superior colliculus. Journal of Neurophysiology 49: 1285–301.CrossRefGoogle ScholarPubMed
Hirsch, J. C. and Burnod, Y. (1987) A synaptically evoked late hyperpolarization in the rat dorsolateral geniculate neurons in vitro. Neuroscience 23: 457–68.CrossRefGoogle ScholarPubMed
Hirsch, J. C., Fourment, A. and Marc, M. E. (1983) Sleep-related variations of membrane potential in the lateral geniculate body relay neurons of the cat. Brain Research 259: 308–12.CrossRefGoogle ScholarPubMed
, N. and Destexhe, A. (2000) Synaptic background activity enhances the responsiveness of neocortical pyramidal neurons. Journal of Neurophysiology 84: 1488–96.CrossRefGoogle ScholarPubMed
Hobin, J. A., Goosens, K. A. and Maren, S. (2003) Context-dependent neuronal activity in the lateral amygdala represents fear memories after extinction. Journal of Neuroscience 23: 8410–16.CrossRefGoogle ScholarPubMed
Hobson, J. A. (1988) The Dreaming Brain. New York: Basic Books.Google Scholar
Hobson, J. A. and Pace-Schott, E. F. (2002) The cognitive neuroscience of sleep: neuronal systems, consciousness and learning. Nature Reviews Neuroscience 3: 679–93.CrossRefGoogle Scholar
Hobson, J. A. and Steriade, M. (1986) Neuronal basis of behavioral state control. In Handbook of Physiology, ed. Mountcastle, V. B. and Bloom, F. E., section 1, vol. IV, pp. 701–823. Bethesda, MD: American Physiological Society.Google Scholar
Hobson, J. A., Pace-Schott, E. and Stickgold, R. (2000) Dreaming and the brain: toward a cognitive neuroscience of conscious states. Brain and Behavioral Sciences 23: 793–842.CrossRefGoogle Scholar
Hofle, N., Paus, T., Reutens, D.et al. (1997) Regional cerebral blood flow changes as a function of delta and spindle activity during slow wave sleep in humans. Journal of Neuroscience 17: 4800–8.CrossRefGoogle ScholarPubMed
Holahan, M. R. and White, N. M. (2002) Conditioned memory modulation, freezing, and avoidance as measures of amygdala-mediated conditioned fear. Neurobiology of Learning and Memory 77: 250–75.CrossRefGoogle ScholarPubMed
Holstege, G. (1990) Subcortical limbic system projections to caudal brainstem and spinal cord. In The Human Nervous System, ed. Paxinos, G., pp. 261–86. New York: Academic Press.Google Scholar
Holstege, G., Bandler, R. and Saper, C. B. (1996) The emotional motor system. Progress in Brain Research 107: 3–6.CrossRefGoogle ScholarPubMed
Homma, Y., Skinner, R. D. and Garcia-Rill, E. (2002) Effects of pedunculopontine nucleus (PPN) stimulation on caudal pontine reticular formation (PnC) neurons in vitro. Journal of Neurophysiology 87: 3033–47.CrossRefGoogle ScholarPubMed
Honda, T. and Semba, K. (1995) An ultrastructural study of cholinergic and non-cholinergic neurons in the laterodorsal and pedunculopontine tegmental nuclei in the rat. Neuroscience 68: 837–53.CrossRefGoogle ScholarPubMed
Hope, B., Michael, G., Knigge, K. and Vincent, S. (1991) Neuronal NADPH diaphorase is a nitric oxide synthase. Proceedings of the National Academy of Sciences USA 88: 2811–14.CrossRefGoogle ScholarPubMed
Hopkins, D. A. and Holstege, G. (1978) Amygdaloid projections to the mesencephalon, pons and medulla oblongata in the cat. Experimental Brain Research 32: 529–47.CrossRefGoogle ScholarPubMed
Hou, Y. P., Manns, I. D. and Jones, B. E. (2002) Immunostaining of cholinergic pontomesencephalic neurons for α1 versus α2 adrenergic receptors suggests different sleep-wake state activities and roles. Neuroscience 114: 517–21.CrossRefGoogle Scholar
Hu, B. (1995) Cellular basis of temporal synaptic signaling: an in vitro electrophysiological study in rat auditory thalamus. Journal of Physiology (London) 483: 167–82.CrossRefGoogle Scholar
Hu, B., Bouhassira, D., Steriade, M. and Deschênes, M. (1988) The blockage of ponto-geniculo-occipital waves in the cat lateral geniculate nucleus by nicotinic antagonists. Brain Research 473: 394–7.CrossRefGoogle ScholarPubMed
Hu, B., Steriade, M. and Deschênes, M. (1989a) The effects of peribrachial stimulation on reticular thalamic neurons: the blockage of spindle waves. Neuroscience 31: 1–12.CrossRefGoogle Scholar
Hu, B., Steriade, M. and Deschênes, M. (1989b) The effects of brainstem peribrachial stimulation on neurons of the lateral geniculate nucleus. Neuroscience 31: 13–24.CrossRefGoogle Scholar
Hu, B., Steriade, M. and Deschênes, M. (1989c) The cellular mechanism of thalamic ponto-geniculo-occipital waves. Neuroscience 31: 25–35.CrossRefGoogle Scholar
Huang, C. C., Hsu, K. S. and Gean, P. W. (1996) Isoproterenol potentiates synaptic transmission primarily by enhancing presynaptic calcium influx via P- and/or Q-type calcium channels in the rat amygdala. Journal of Neuroscience 16: 1026–33.CrossRefGoogle ScholarPubMed
Huang, Y. Y. and Kandel, E. R. (1998) Postsynaptic induction and PKA-dependent expression of LTP in the lateral amygdala. Neuron 21: 169–78.CrossRefGoogle ScholarPubMed
Hubel, D. H. and Wiesel, T. N. (1961) Integrative action in the cat's lateral geniculate body. Journal of Physiology (London) 155: 385–98.CrossRefGoogle ScholarPubMed
Huber, R., Ghilardi, M. F., Massimini, M. and Tononi, G. (2004) Local sleep and learning. Nature 430: 78–81.CrossRefGoogle Scholar
Hugues, S., Deschaux, O. and Garcia, R. (2004) Postextinction infusion of a mitogen-activated protein kinase inhibitor into the medial prefrontal cortex impairs memory of the extinction of conditioned fear. Learning and Memory 11: 540–3.CrossRefGoogle ScholarPubMed
Hughes, S. W., Cope, D. W., Toth, T. I., Williams, S. R. and Crunelli, V. (1999) All thalamocortical neurones possess a T-type Ca2+ ‘window’ current that enables the expression of bistability-mediated activities. Journal of Physiology (London) 517: 805–15.CrossRefGoogle ScholarPubMed
Hughes, S. W., Blethyn, K. L., Cope, D. W. and Crunelli, V. (2002a) Properties and origin of spikelets in thalamocortical neurones in vitro. Neuroscience 110: 395–401.CrossRefGoogle Scholar
Hughes, S. W., Cope, D. W., Blethyn, K. L. and Crunelli, V. (2002b) Cellular mechanisms of the slow (<1 Hz) oscillation in thalamocortical neurons in vitro. Neuron 33: 947–58.CrossRefGoogle Scholar
Hughes, S. W., Lörincz, M., Cope, D. W.et al. (2004) Synchronized oscillations at alpha and theta frequencies in the lateral geniculate nucleus. Neuron 42: 253–68.CrossRefGoogle ScholarPubMed
Huguenard, J. R. (1996) Low-threshold calcium currents in central nervous system neurons. Annual Review of Physiology 58: 329–48.CrossRefGoogle ScholarPubMed
Huguenard, J. R. and Prince, D. A. (1992) A novel T-type current underlies prolonged Ca2+-dependent burst firing in GABAergic neurons of rat thalamic reticular nucleus. Journal of Neuroscience 12: 3804–17.CrossRefGoogle Scholar
Huguenard, J. R. and Prince, D. A. (1994) Intrathalamic rhythmicity studied in vitro: nominal T current modulation causes robust anti-oscillatory effects. Journal of Neuroscience 14: 5485–502.CrossRefGoogle Scholar
Hugues, S., Deschaux, O. and Garcia, R. (2004) Postextinction infusion of a mitogen-activated protein kinase inhibitor into the medial prefrontal cortex impairs memory of the extinction of conditioned fear. Learning and Memory 11: 540–3.CrossRefGoogle ScholarPubMed
Humeau, Y., Shaban, H., Bissière, S. and Lüthi, A. (2003) Presynaptic induction of heterosynaptic associative plasticity in the mammalian brain. Nature 426: 841–5.CrossRefGoogle ScholarPubMed
Humeau, Y., Herry, C., Kemp, N.et al. (2005) Dendritic spine heterogeneity determines afferent-specific Hebbian plasticity in the amygdala. Neuron 45: 119–31.CrossRefGoogle ScholarPubMed
Humphrey, T. (1936) The telencephalon of the bat. I. The noncortical nuclear masses and certain pertinent fiber connections. Journal of Comparative Neurology 65: 603–711.CrossRefGoogle Scholar
Hunt, C. A., Pang, D. Z. and Jones, E. G. (1991) Distribution and density of GABA cells in intralaminar and adjacent nuclei of monkey thalamus. Neuroscience 43: 185–96.CrossRefGoogle ScholarPubMed
Hurley, K. M., Herbert, H., Moga, M. M. and Saper, C. B. (1991) Efferent projections of the infralimbic cortex of the rat. Journal of Comparative Neurology 308: 249–76.CrossRefGoogle ScholarPubMed
Hutcheon, B., Miura, R. M. and Puil, E. (1996) Subthreshold membrane resonance in neocortical neurons. Journal of Neurophysiology 76: 683–97.CrossRefGoogle ScholarPubMed
Inglis, W. L. and Semba, K. (1996) Colocalization of ionotropic glutamate receptor subunits with NADPH-diaphorase-containing neurons in the rat mesopontine tegmentum. Journal of Comparative Neurology 368: 17–32.3.0.CO;2-N>CrossRefGoogle ScholarPubMed
Ingvar, D. H., Sjölund, B. and Ardo, A. (1976) Correlation between ECG frequency, cerebral oxygen uptake and blood flow. Electroencephalography and Clinical Neurophysiology 41: 268–76.CrossRefGoogle Scholar
Insausti, R., Amaral, D. G. and Cowan, W. M. (1987) The entorhinal cortex of the monkey: III. Subcortical afferents. Journal of Comparative Neurology 264: 396–408.CrossRefGoogle ScholarPubMed
Introini-Collison, I. B., Miyazaki, B. and McGaugh, J. L. (1991) Involvement of the amygdala in the memory-enhancing effects of clenbuterol. Psychopharmacology 104: 541–4.CrossRefGoogle ScholarPubMed
Inubushi, S., Kobayashi, T., Oshima, T. and Torii, S. (1978a) Intracellular recordings from motor cortex during EEG arousal in unanesthetized brain preparations of the cat. Japanese Journal of Physiology 28: 669–88.CrossRefGoogle Scholar
Inubushi, S., Kobayashi, T., Oshima, T. and Torii, S. (1978b) An intracellular analysis of EEG arousal in cat motor cortex. Japanese Journal of Physiology 28: 689–708.CrossRefGoogle Scholar
Irle, E. and Markowitsch, H. J. (1984) Basal forebrain efferents reach the whole cerebral cortex of the cat. Brain Research Bulletin 12: 493–512.CrossRefGoogle ScholarPubMed
Ito, K. and McCarley, R. W. (1984) Alterations in membrane potential and excitability of cat medial pontine reticular formation neurons during changes in naturally occurring sleep-wake states. Brain Research 292: 169–75.CrossRefGoogle ScholarPubMed
Iwata, J., LeDoux, J. E., Meeley, M. P., Arneric, S. and Reis, D. J. (1986) Intrinsic neurons in the amygdaloid field projected to by the medial geniculate body mediate emotional responses conditioned to acoustic stimuli. Brain Research 383: 195–214.CrossRefGoogle ScholarPubMed
Izquierdo, I. and Medina, J. H. (1993) Role of the amygdala, hippocampus and entorhinal cortex in memory consolidation and expression. Brazilian Journal of Medical and Biological Research 26: 573–89.Google ScholarPubMed
Jacobs, B. L. and McGinty, D. J. (1971) Amygdala unit activity during sleep and waking. Experimental Neurology 33: 1–15.CrossRefGoogle ScholarPubMed
Jahnsen, H. and Llinás, R. (1984a) Electrophysiological properties of guinea-pig thalamic neurones: an in vitro study. Journal of Physiology (London) 349: 205–26.CrossRefGoogle Scholar
Jahnsen, H. and Llinás, R. (1984b) Ionic basis for electroresponsiveness and oscillatory properties of guinea-pig thalamic neurones in vitro. Journal of Physiology (London) 349: 227–47.CrossRefGoogle Scholar
Jasper, H. H. (1949) Diffuse projection systems: the integrative action of the thalamic reticular system. Electroencephalography and Clinical Neurophysiology 1: 405–20.CrossRefGoogle ScholarPubMed
Jasper, H. H. (1958) Recent advances in our understanding of ascending activities of the reticular system. In Reticular Formation of the Brain, ed. Jasper, H. H., Proctor, L. D., Knighton, R. S., Noshay, W. C. and Costello, R. T., pp. 319–31. Boston, Toronto: Little-Brown.Google Scholar
Jay, T. M., Burette, F. and Laroche, S. (1996) Plasticity of the hippocampal-prefrontal cortex synapses. Journal of Physiology (Paris) 90: 361–6.CrossRefGoogle ScholarPubMed
Jeanmonod, D., Magnin, M. and Morel, A. (1996) Low-threshold calcium spike-bursts in the human thalamus: common physiopathology for sensory, motor and limbic positive symptoms. Brain 119: 363–75.CrossRefGoogle ScholarPubMed
Jenkins, J. and Dallenback, K. (1924) Oblivescence during sleep and waking. American Journal of Psychology 35: 605–12.CrossRefGoogle Scholar
Johnston, J. B. (1923) Further contributions to the study of the evolution of the forebrain. Journal of Comparative Neurology 35: 337–481.CrossRefGoogle Scholar
Jolkkonen, E. and Pitkänen, A. (1998) Intrinsic connections of the rat amygdaloid complex: projections originating in the central nucleus. Journal of Comparative Neurology 395: 53–72.3.0.CO;2-G>CrossRefGoogle ScholarPubMed
Jolkkonen, E., Miettinen, R., Pikkarainen, M. and Pitkänen, A. (2002) Projections from the amygdaloid complex to the magnocellular cholinergic basal forebrain in rat. Neuroscience 111: 133–49.CrossRefGoogle ScholarPubMed
Jones, B. E. (2003) Arousal systems. Frontiers in Bioscience 8: s438–51.CrossRefGoogle ScholarPubMed
Jones, B. E. (2004) Activity, modulation and role of basal forebrain cholinergic neurons innervating the cerebral cortex. In Acetylcholine in the Cerebral Cortex (Progress in Brain Research, vol. 145), ed. Descarries, L., Krnjević, K. and Steriade, M., pp. 157–69. Amsterdam: Elsevier.Google Scholar
Jones, B. E. and Beaudet, A. (1987) Distribution of acetylcholine and catecholamine neurons in the cat brain stem studied by choline acetyltransferase and tyrosine hydroxylase immunohistochemistry. Journal of Comparative Neurology 261: 15–32.CrossRefGoogle Scholar
Jones, B. E., and Cuello, A. C. (1989) Afferents to the basal forebrain cholinergic cell area from pontomesencephalic — catecholamine, serotonin, and acetylcholine — neurons. Neuroscience 31: 37–61.CrossRefGoogle ScholarPubMed
Jones, B. E. and Moore, R. Y. (1977) Ascending projections of the locus coeruleus in the rat. II. Autoradiographic study. Brain Research 127: 25–53.CrossRefGoogle ScholarPubMed
Jones, E. G. (1975) Varieties and distribution of non-pyramidal cells in the somatic sensory cortex of the squirrel monkey. Journal of Comparative Neurology 160: 205–68.CrossRefGoogle ScholarPubMed
Jones, E. G. (1984) Laminar distribution of cortical efferent cells. In Cerebral Cortex, vol. 1 (Cellular Components of the Cerebral Cortex), ed. Peters, A. and Jones, E. G., pp. 521–33. New York: Plenum.Google Scholar
Jones, E. G. (1985) The Thalamus. New York: Plenum.CrossRefGoogle Scholar
Jones, E. G. (1991) Cellular organization in the primate postcentral gyrus. In Information Processing in the Somatosensory System, ed. Franzen, O. and Westman, J., pp. 95–107. New York: Macmillan.CrossRefGoogle Scholar
Jones, E. G. (1997). A description of the human thalamus. In Thalamus, vol. 2 (Experimental and Clinical Aspects), ed. Steriade, M., Jones, E. G. and McCormick, D. A., pp. 425–99. Amsterdam: Elsevier.Google Scholar
Jones, E. G. (2000) Cortical and subcortical contributions to activity-dependent plasticity in primate somatosensory cortex. Annual Reviews of Neuroscience 23: 1–37.CrossRefGoogle ScholarPubMed
Jones, E. G. (2002) Thalamic circuitry and thalamocortical synchrony. Philosophical Transactions of the Royal Society of LondonB357: 1659–73.CrossRefGoogle ScholarPubMed
Jones, E. G. and Pons, T. P. (1998) Thalamic and brainstem contributions to large-scale plasticity of primate somatosensory cortex. Science 282: 1121–25.CrossRefGoogle ScholarPubMed
Jones, E. G. and Powell, T. P. S. (1970) An anatomical study of converging sensory pathways within the cerebral cortex of the monkey. Brain 93: 793–820.CrossRefGoogle ScholarPubMed
Jones, R. S. G. (1990) Intrinsic properties of neurones in layer II of the rat entorhinal cortex in vitro. Journal of Physiology (London) 425: 86P.Google Scholar
Jones, R. S. G. (1994) Synaptic and intrinsic properties of neurons of origin of the perforant path in layer II of the rat entorhinal cortex in vitro. Hippocampus 4: 335–53.CrossRefGoogle ScholarPubMed
Jones, R. S. G. and Bühl, E. H. (1993) Basket-like interneurones in layer II of the entorhinal cortex exhibit a powerful NMDA-mediated synaptic excitation. Neuroscience Letters 149: 35–9.CrossRefGoogle ScholarPubMed
Jongen-Rêlo, A. L. and Amaral, D. G. (1998) Evidence for a GABAergic projection from the central nucleus of the amygdala to the brainstem of the macaque monkey: a combined retrograde tracing and in situ hybridization study. European Journal of Neuroscience 10: 2924–33.CrossRefGoogle ScholarPubMed
Jouvet, M. (1965) Paradoxical sleep – a study of its nature and mechanisms. Progress in Brain Research 18: 20–57.CrossRefGoogle ScholarPubMed
Jouvet, M. (1972) The role of monoamines and acetylcholine-containing neurons in the regulation of the sleep-waking cycle. Ergebnisse der Physiologie 64: 166–307.Google ScholarPubMed
Jouvet, M. and Delorme, J. F. (1965) Locus coeruleus et sommeil paradoxal. Comptes Rendus de la Société de Biologie de Paris 159: 895–9.Google Scholar
Jouvet, M. and Michel, F. (1959) Corrélations electromyographiques du sommeil chez le chat décortiqué et mésencephalique chronique. Comptes Rendus de la Société de Biologie de Paris 153: 422–5.Google Scholar
Kamal, A. M. and Tömböl, T. (1975) Golgi studies on the amygdaloid nuclei of the cat. Journal für Hirnforschung 16: 175–201.Google ScholarPubMed
Kammermeier, P. J. and Jones, S. W. (1997) High-voltage-activated calcium currents in neurons acutely isolated from the ventrobasal nucleus of the rat thalamus. Journal of Neurophysiology 77: 465–75.CrossRefGoogle ScholarPubMed
Kamondi, A., Williams, J. A., Hutcheon, B. and Reiner, P. B. (1992) Membrane properties of mesopontine cholinergic neurons studied with the whole-cell patch-clamp technique: implications for behavioral state control. Journal of Neurophysiology 68: 1359–72.CrossRefGoogle ScholarPubMed
Kang, Y. and Kayano, F. (1994) Electrophysiological and morphological characteristics of layer VI pyramidal cells in the cat motor cortex. Journal of Neurophysiology 72: 578–91.CrossRefGoogle ScholarPubMed
Kang, Y. and Kitai, S. T. (1990) Electrophysiological properties of pedunculopontine neurons and their postsynaptic responses following stimulation of substantia nigra reticulata. Brain Research 535: 79–95.CrossRefGoogle ScholarPubMed
Kapp, B. S., Frysinger, R. C., Gallagher, M. and Haselton, J. R. (1979) Amygdala central nucleus lesions: Effects on heart rate conditioning in the rabbit. Physiology and Behavior 23: 1109–17.CrossRefGoogle Scholar
Kapp, B. S., Supple, W. J. and Whalen, P. J. (1994) Effects of electrical stimulation of the amygdaloid central nucleus on neocortical arousal in the rabbit. Behavioral Neuroscience 108: 81–93.CrossRefGoogle ScholarPubMed
Kasamatsu, T. (1976) Visual cortical neurons influenced by the oculomotor input: characterization of their receptive field properties. Brain Research 113: 271–92.CrossRefGoogle ScholarPubMed
Kasper, E. M., Larkman, A. U., Lübke, J. and Blakemore, C. (1994a) Pyramidal neurons in layer 5 of the rat visual cortex. I. Correlation among cell morphology, intrinsic electrophysiological properties, and axon targets. Journal of Comparative Neurology 339: 459–74.CrossRefGoogle Scholar
Kasper, E. M., Lübke, J., Larkman, A. U. and Blakemore, C. (1994b) Pyramidal neurons in layer 5 of the rat visual cortex. III. Differential maturation of axon targeting, dendritic morphology, and electrophysiological properties. Journal of Comparative Neurology 339: 495–518.CrossRefGoogle Scholar
Kato, N. (1990) Cortico-thalamo-cortical projection between visual cortices. Brain Research 509: 150–2.CrossRefGoogle ScholarPubMed
Kawaguchi, Y. (1993) Groupings of nonpyramidal and pyramidal cells with specific physiological and morphological characteristics in rat frontal cortex. Journal of Neurophysiology 69: 416–31.CrossRefGoogle ScholarPubMed
Kawaguchi, Y. (2001) Distinct firing patterns of neuronal subtypes in cortical synchronized activities. Journal of Neuroscience 21: 7261–72.CrossRefGoogle ScholarPubMed
Kawaguchi, Y. and Kubota, Y. (1993) Correlation of physiological subgroupings of nonpyramidal cells with parvalbumine- and calbindinD28k-immunoreactive neurons in layer V of rat frontal cortex. Journal of Neurophysiology 70: 387–96.CrossRefGoogle Scholar
Kawaguchi, Y. and Kubota, Y. (1997) GABAergic cell subtypes and their synaptic connections in rat frontal cortex. Cerebral Cortex 7: 476–86.CrossRefGoogle ScholarPubMed
Kayama, Y., Ohta, M. and Jodo, E. (1992) Firing of ‘possibly’ cholinergic neurons in the rat laterodorsal tegmental nucleus during sleep and wakefulness. Brain Research 569: 210–20.CrossRefGoogle ScholarPubMed
Kemp, N. and Bashir, Z. I. (2001) Long-term depression: a cascade of induction and expression mechanisms. Progress in Neurobiology 65: 339–65.CrossRefGoogle ScholarPubMed
Kemppainen, S. and Pitkänen, A. (2000) Distribution of parvalbumin, calretinin, and calbindin-D28k immunoreactivity in the rat amygdaloid complex and colocalization with gamma-aminobutyric acid. Journal of Comparative Neurology 426: 441–67.3.0.CO;2-7>CrossRefGoogle Scholar
Kettenmann, H. and Schachner, M. (1985) Pharmacological properties of γ-aminobutyric acid-, glutamate-, and aspartate-induced depolarizations in cultured astrocytes, Journal of Neuroscience 5: 3295–301.CrossRefGoogle ScholarPubMed
Khateb, A., Mühlethaler, M., Alonso, A.et al. (1992) Cholinergic nucleus basalis neurons display the capacity for rhythmic bursting activity mediated by low-threshold calcium spikes. Neuroscience 51: 489–94.CrossRefGoogle ScholarPubMed
Khateb, A., Fort, P., Serafin, M., Jones, B. E. and Mühlethaler, M. (1995a) Rhythmical bursts induced by NMDA in guinea-pig cholinergic nucleus basalis neurones in vitro. Journal of Physiology (London) 487: 623–38.CrossRefGoogle Scholar
Khateb, A., Fort, P., Pegna, A., Jones, B. E. and Mühlethaler, M. (1995b) Cholinergic nucleus basalis neurons are excited by histamine in vitro. Neuroscience 69: 495–506.CrossRefGoogle Scholar
Khateb, A., Fort, P., Williams, S.et al. (1997) Modulation of cholinergic nucleus basalis neurons by acetylcholine and N-methyl-D-aspartate. Neuroscience 81: 47–55.CrossRefGoogle ScholarPubMed
Killcross, S., Robbins, T. W. and Everitt, B. J. (1997) Different types of fear-conditioned behaviour mediated by separate nuclei within amygdala. Nature 388: 377–80.CrossRefGoogle ScholarPubMed
Kim, J. J. and Fanselow, M. S. (1992) Modality-specific retrograde amnesia of fear. Science 256: 675–7.CrossRefGoogle Scholar
Kim, M. J., Chun, S. K., Kim, Y. B., Mook-Jung, I. and Jung, M. W. (2003) Long-term potentiation in visual cortical projections to the medial prefrontal cortex of the rat. Neuroscience 120: 283–9.CrossRefGoogle ScholarPubMed
Kim, U. and McCormick, D. A. (1998) Functional and ionic properties of a slow afterhyperpolarization in ferret perigeniculate neurones in vitro. Journal of Neurophysiology 80: 1222–35.CrossRefGoogle Scholar
Kim, U., Bal, T. and McCormick, D. A. (1995) Spindle waves are propagating synchronized oscillations in the ferret LGNd in vitro. Journal of Neurophysiology 74: 1301–23.CrossRefGoogle ScholarPubMed
Kinney, H. C., Korein, J., Panigrahy, A., Dikkes, P. and Goode, R. (1994) Neuropathological findings in the brain of Karen Ann Quinlan. New England Journal of Medicine 330: 1469–75.CrossRefGoogle ScholarPubMed
Kinomura, S., Larsson, J., Gulyás, B. and Roland, P. (1996) Activation by attention of the human reticular formation and thalamic intralaminar nuclei. Science 271: 512–15.CrossRefGoogle ScholarPubMed
Kisvárday, Z. F., Beaulieu, C. and Eysel, U. T. (1993) Network of GABAergic large basket cells in cat visual cortex (area 18): implication for lateral disinhibition. Journal of Comparative Neurology 327: 398–415.CrossRefGoogle ScholarPubMed
Kitsikis, A. and Steriade, M. (1981) Immediate behavioral effects of kainic acid injections into the midbrain reticular core. Behavioral Brain Research 3: 361–80.CrossRefGoogle ScholarPubMed
Kleitman, N. (1963) Sleep and Wakefulness. Chicago: University of Chicago Press.Google Scholar
Klink, R. and Alonso, A. (1993) Ionic mechanisms for the subthreshold oscillations and differential electroresponsiveness of medial entorhinal cortex layer II neurons. Journal of Neurophysiology 70: 149–57.CrossRefGoogle ScholarPubMed
Kobayashi, Y., Inoue, Y., Yamamoto, M., Isa, T. and Aizawa, H. (2001) Contribution of pedunculopontine tegmental nucleus neurons to performance of visually guided saccade tasks in monkeys. Journal of Neurophysiology 88: 715–31.CrossRefGoogle Scholar
Koch, C. (1998) The neuroanatomy of visual consciousness. In Consciousness: At the Frontiers of Neuroscience (Advances in Neurology, vol. 77), ed. Jasper, H. H., Descarries, L., Castelucci, V. F. and Rossignol, S., pp. 229–41. Philadelphia, PA: Lippincott-Raven.Google Scholar
Köhler, C. and Steinbusch, H. (1982) Identification of serotonin and non-serotonin-containing neurons of the mid-brain raphe projecting to the entorhinal area and the hippocampal formation. A combined immunohistochemical and fluorescent retrograde tracing study in the rat brain. Neuroscience 7: 951–75.CrossRefGoogle ScholarPubMed
Köhler, C., Chan-Palay, V., Haglund, L. and Steinbusch, H. (1980) Immunohistochemical localization of serotonin nerve terminals in the lateral entorhinal cortex of the rat: demonstration of two separate patterns of innervation from the midbrain raphe. Anatomy and Embryology 160: 121–9.CrossRefGoogle ScholarPubMed
Köhler, C., Chan-Palay, V. and Steinbusch, H. (1981) The distribution and orientation of serotonin fibers in the entorhinal and other retrohippocampal areas. An immunohistochemical study with anti-serotonin antibodies in the rat brain. Anatomy and Embryology 161: 237–64.CrossRefGoogle Scholar
Köhler, C., Swanson, L. W., Haglund, L. and Wu, J. Y. (1985a) The cytoarchitecture, histochemistry and projections of the tuberomammillary nucleus in the rat. Neuroscience 16: 85–110.CrossRefGoogle Scholar
Köhler, C., Wu, J. Y. and Chan-Palay, V. (1985b) Neurons and terminals in the retrohippocampal region in he rat's brain identified by anti-gamma-aminobutyric acid and anti-glutamic acid decarboxylase immunocytochemistry. Anatomy and Embryology (Berlin) 173: 35–44.CrossRefGoogle Scholar
Kohlmeier, K. A. and Reiner, P. B. (1999) Noradrenaline excites non-cholinergic laterodorsal tegmental neurons via two distinct mechanisms. Neuroscience 93: 619–30.CrossRefGoogle ScholarPubMed
Koo, J. W., Han, J. S. and Kim, J. J. (2004) Selective neurotoxic lesions of basolateral and central nuclei of the amygdala produce differential effects on fear conditioning. Journal of Neuroscience 24: 7654–62.CrossRefGoogle ScholarPubMed
Kordower, J. H., Bartus, R. T., Marciano, F. F. and Gash, D. M. (1989) Telencephalic cholinergic system of the new world monkey (Cebus apella): Morphological and cytoarchitectonic assessment and analysis of the projection to the amygdala. Journal of Comparative Neurology 279: 528–45.CrossRefGoogle ScholarPubMed
Kosaka, T., Kosaka, K., Hataguchi, Y.et al. (1987) Catecholaminergic neurons containing GABA-like and/or glutamic acid decarboxylase-like immunoreactivities in various brain regions of the rat. Experimental Brain Research 66: 191–210.CrossRefGoogle ScholarPubMed
Kosofsky, B. E. and Molliver, M. E. (1987) The serotoninergic innervation of cerebral cortex: different classes of axon terminals arise from dorsal and median raphe nuclei. Synapse 1: 153–68.CrossRefGoogle ScholarPubMed
Krettek, J. E. and Price, J. L. (1977a) Projections from the amygdaloid complex to the cerebral cortex and thalamus in the rat and cat. Journal of Comparative Neurology 172: 687–722.CrossRefGoogle Scholar
Krettek, J. E. and Price, J. L. (1977b) Projections from the amygdaloid complex and adjacent olfactory structures to the entorhinal cortex and to the subiculum in the rat and cat. Journal of Comparative Neurology 172: 723–52.CrossRefGoogle Scholar
Krettek, J. E. and Price, J. L. (1978a) Amygdaloid projections to subcortical structures within the basal forebrain and brainstem in the rat and cat. Journal of Comparative Neurology 178: 225–54.CrossRefGoogle Scholar
Krettek, J. E. and Price, J. L. (1978b) A description of the amygdaloid complex in the rat and cat with observations on intra-amygdaloid axonal connections. Journal of Comparative Neurology 178: 255–80.CrossRefGoogle Scholar
Kroner, S., Rosenkranz, J. A., Grace, A. A. and Barrionuevo, G. (2005) Dopamine modulates excitability of basolateral amygdala neurons in vitro. Journal of Neurophysiology 93: 1598–610.CrossRefGoogle ScholarPubMed
Kudo, M. and Niimi, K. (1980) Ascending projections of the inferior colliculus in the cat: an autoradiographic study. Journal of Comparative Neurology 191: 545–56.CrossRefGoogle Scholar
Kudo, M., Itoh, K., Kawamura, S. and Mizuno, N. (1983) Direct projections to the pretectum and the midbrain reticular formation from auditory relay nuclei in the lower brainstem of the cat. Brain Research 288: 13–19.CrossRefGoogle ScholarPubMed
Kudo, M., Tashiro, T., Higo, S., Matsuyama, T. and Kawamura, S. (1984) Ascending projections from the nucleus of the brachium of the inferior colliculus in the cat. Experimental Brain Research 54: 203–11.CrossRefGoogle ScholarPubMed
Kumar, S. S. and Huguenard, J. R. (2001) Properties of excitatory synaptic connections mediated by the corpus callosum in the developing neocortex. Journal of Neurophysiology 86: 2973–85.CrossRefGoogle Scholar
Kuroda, M. and Price, J. L. (1991) Synaptic organization of projections from basal forebrain structures to the mediodorsal thalamic nucleus of the rat. Journal of Comparative Neurology 303: 513–33.CrossRefGoogle ScholarPubMed
LaBar, K. S., LeDoux, J. E., Spencer, D. D. and Phelps, E. A. (1995) Impaired fear conditioning following unilateral temporal lobectomy in humans. Journal of Neuroscience 15: 6846–55.CrossRefGoogle ScholarPubMed
Lamprecht, R., Farb, C. R. and LeDoux, J. E. (2002) Fear memory formation involves p190 RhoGAP and ROCK proteins through a GRB2-mediated complex. Neuron 36: 727–38.CrossRefGoogle ScholarPubMed
Lancel, M., Riezen, H., and Glatt, A. (1992) The time course of sigma activity and slow-wave activity during NREMS in cortical and thalamic EEG of the cat during baseline and after 12 hours of wakefulness. Brain Research 596: 285–95.CrossRefGoogle ScholarPubMed
Landisman, C. E., Long, M. A., Beierlein, M.et al. (2002) Electrical synapses in the thalamic reticular nucleus. Journal of Neuroscience 22: 1002–9.CrossRefGoogle ScholarPubMed
Lang, E. J. and Paré, D. (1997a) Similar inhibitory processes dominate the responses of cat lateral amygdaloid projection neurons to their various afferents. Journal of Neurophysiology 77: 341–52.CrossRefGoogle Scholar
Lang, E. J. and Paré, D. (1997b) Synaptic and synaptically activated intrinsic conductances underlie inhibitory potentials in cat lateral amygdaloid projection neurons in vivo. Journal of Neurophysiology 77: 353–63.CrossRefGoogle Scholar
Lang, E. J. and Paré, D. (1998) Synaptic responsiveness of interneurons of the cat lateral amygdaloid nucleus. Neuroscience 83: 877–89.CrossRefGoogle ScholarPubMed
Larkum, M. E., Zhu, J. J. and Sakman, B. (1999) A new cellular mechanism for coupling inputs arriving at different cortical layers. Nature 398: 338–41.CrossRefGoogle ScholarPubMed
Laroche, S., Jay, T. M. and Thierry, A. M. (1990) Long-term potentiation in the prefrontal cortex following stimulation of the hippocampal CA1/subicular region. Neuroscience Letters 114: 184–90.CrossRefGoogle ScholarPubMed
Laroche, S., Doyère, V., Redini-Del Negro, C. and Burette, F. (1995) Neural mechanisms of associative memory: role of long-term potentiation. In Brain and Memory: Modulation and Mediation of Neuroplasticity, ed. McGaugh, J. L., Weinberger, N. M. and Lynch, G., pp. 277–302. New York: Oxford University Press.CrossRefGoogle Scholar
Larson, J. and Lynch, G. (1986) Role of N-methyl-D aspartate receptors in the induction of synaptic potentiation by burst stimulation patterned after the hippocampal theta rhythm. Brain Research 441: 111–18.CrossRefGoogle Scholar
Larson, J., Wong, D. and Lynch, G. (1986) Patterned stimulation at the theta frequency is optimal for the induction of hippocampal long-term potentiation. Brain Research 368: 347–50.CrossRefGoogle ScholarPubMed
Lavin, A. and Grace, A. A. (1994) Modulation of dorsal thalamic cell activity by the ventral pallidum: its role in the regulation of thalamocortical activity by the basal ganglia. Synapse 18: 104–27.CrossRefGoogle ScholarPubMed
Lavoie, B. and Parent, A. (1994) Pedunculopontine nucleus in the squirrel monkey: distribution of cholinergic and monoaminergic neurons in the mesopontine tegmentum with evidence for the presence of glutamate in cholinergic neurons. Journal of Comparative Neurology 344: 190–209.CrossRefGoogle ScholarPubMed
Law-Tho, D., Desce, J. M. and Crépel, F. (1995) Dopamine favours the emergence of long-term depression versus long-term potentiation in slices of rat prefrontal cortex. Neuroscience Letters 188: 125–8.CrossRefGoogle ScholarPubMed
LeDoux, J. E. (2000a) The amygdala and emotion: a view through fear. In The Amygdala: A Functional Analysis, ed. Aggleton, J. P., pp. 289–310. Oxford: Oxford University Press.Google Scholar
LeDoux, J. E. (2000b) Emotion circuits in the brain. Annual Review of Neuroscience 23: 155–84.CrossRefGoogle Scholar
LeDoux, J. E., Sakaguchi, A. and Reis, D. J. (1984) Subcortical efferent projections of the medial geniculate nucleus mediate emotional responses conditioned to acoustic stimuli. Journal of Neuroscience 4: 683–98.CrossRefGoogle ScholarPubMed
LeDoux, J. E., Ruggiero, D. A., Forest, R., Stornetta, R. and Reis, D. J. (1987) Topographic organization of convergent projections to the thalamus from the inferior colliculus and spinal cord in the rat. Journal of Comparative Neurology 264: 123–46.CrossRefGoogle ScholarPubMed
LeDoux, J. E., Iwata, J., Cicchetti, P. and Reis, D. J. (1988) Different projections of the central amygdaloid nucleus mediate autonomic and behavioral correlates of conditioned fear. Journal of Neuroscience 8: 2517–29.CrossRefGoogle ScholarPubMed
LeDoux, J. E., Cicchetti, P., Xagoraris, A. and Romanski, L. M. (1990a) The lateral amygdaloid nucleus: sensory interface of the amygdala in fear conditioning. Journal of Neuroscience 10: 1062–9.CrossRefGoogle Scholar
LeDoux, J. E., Farb, C. and Ruggiero, D. A. (1990b) Topographic organization of neurons in the acoustic thalamus that project to the amygdala. Journal of Neuroscience 10: 1043–54.CrossRefGoogle Scholar
Lee, H. and Kim, J. J. (1998) Amygdalar NMDA receptors are critical for new fear learning in previously fear-conditioned rats. Journal of Neuroscience 18: 8444–54.CrossRefGoogle ScholarPubMed
Lee, H. K., Barbarosie, M., Kameyama, K., Bear, M. F. and Huganir, R. L. (2000) Regulation of distinct AMPA receptor phosphorylation sites during bidirectional synaptic plasticity. Nature 405: 955–9.CrossRefGoogle ScholarPubMed
Lee, K. H. and McCormick, D. A. (1995) Acetylcholine excites GABAergic neurons of the ferret perigeniculate nucleus through nicotinic receptors. Journal of Neurophysiology 73: 2123–8.CrossRefGoogle ScholarPubMed
Lee, K. and McCormick, D. A. (1996) Abolition of spindle oscillations by serotonin and norepinephrine in the ferret lateral geniculate and perigeniculate nuclei in vitro. Neuron 17: 309–21.CrossRefGoogle ScholarPubMed
Lee, M. G., Hassani, O. K., Alonso, A. and Jones, B. E. (2005) Cholinergic basal forebrain neurons burst with theta during waking and paradoxical sleep. Journal of Neuroscience 25: 4365–9.CrossRefGoogle ScholarPubMed
Lee, M. S., Friedberg, M. H. and Ebner, F. F. (1994) The role of GABA-mediated inhibition in the rat ventral posterior medial thalamus. I. Assessment of receptive field changes following thalamic reticular nucleus lesions. Journal of Neurophysiology 71: 1702–15.CrossRefGoogle ScholarPubMed
Lenz, F. A. and Dougherty, P. M. (1997) Pain processing in the human thalamus. In Thalamus, vol. 2 (Experimental and Clinical Aspects), ed. Steriade, M., Jones, E. G. and McCormick, D. A., pp. 617–51. Amsterdam: Elsevier.
Leonard, B. W., Amaral, D. G., Squire, L. R. and Zola-Morgan, S. (1995) Transient memory impairment in monkeys with bilateral lesions of the entorhinal cortex. Journal of Neuroscience 15: 5367–659.CrossRefGoogle ScholarPubMed
Leonard, C. S. and Llinás, R. R. (1990) Electrophysiology of mammalian pedunculopontine and laterodorsal tegmental neurons in vitro: implications for the control of REM sleep. In Brain Cholinergic Systems, ed. Steriade, M. and Biesold, D., pp. 205–23. Oxford: Oxford University Press.
Leonard, C. S. and Llinás, R. R. (1994) Serotonergic and cholinergic inhibition of mesopontine cholinergic neurons controlling REM sleep: an in vitro electrophysiological study. Neuroscience 59: 309–30.CrossRefGoogle Scholar
Leonard, C. S., Kerman, I., Blaha, G., Taveras, E., and Taylor, B. (1995a) Interdigitation of nitric oxyde synthase-, tyrosine hydroxylase- and serotonin-containing neurons in and around the laterodorsal and pedunculopontine tegmental nuclei of the guinea pig. Journal of Comparative Neurology 362: 411–32.CrossRefGoogle Scholar
Leonard, C. S., Rao, S. and Sanchez, R. M. (1995b) Patterns of neuromodulation and the nitric oxide signaling pathway in mesopontine cholinergic neurons. Seminars in the Neurosciences 7: 319–28.CrossRefGoogle Scholar
Leonard, C. S., Rao, S. R. and Inoue, T. (2000) Serotonergic inhibition of action potential evoked calcium transients in NOS-containing mesopontine cholinergic neurons. Journal of Neurophysiology 84: 1558–72.CrossRefGoogle ScholarPubMed
Leopold, D. A., Murayama, Y. and Logothetis, N. K. (2003) Very slow activity fluctuations in monkey visual cortex: implications for functional brain imaging. Cerebral Cortex 13: 422–33.CrossRefGoogle ScholarPubMed
Leresche, N., Jassik-Gerschenfeld, D., Haby, M., Soltesz, I. and Crunelli, V. (1990) Pacemaker-like and other types of spontaneous membrane potential oscillations of thalamocortical cells. Neuroscience Letters 113: 72–7.CrossRefGoogle ScholarPubMed
Leresche, N., Lightowler, S., Soltesz, I., Jassik-Gerschenfeld, D. and Crunelli, V. (1991) Low-frequency oscillatory activities intrinsic to rat and cat thalamocortical cells. Journal of Physiology (London) 441: 155–74.CrossRefGoogle ScholarPubMed
Leung, L. S. (1984) Model of gradual phase shift of theta rhythm in the rat. Journal of Neurophysiology 52: 1051–65.CrossRefGoogle ScholarPubMed
LeVay, S. (1973) Synaptic patterns in the visual cortex of the cat and monkey. Electron microscopy of Golgi preparations. Journal of Comparative Neurology 150: 53–86.CrossRefGoogle ScholarPubMed
Lévesque, M., Charara, A., Gagnon, S., Parent, A. and Deschênes, M. (1996) Corticostriatal projections from layer V cells in rat are collaterals of long-range corticofugal axons. Brain Research 709: 311–15.CrossRefGoogle Scholar
Levitt, P. and Moore, R. Y. (1978) Noradrenaline neuron innervation of the neocortex in the rat. Brain Research 139: 219–31.CrossRefGoogle ScholarPubMed
Lhermitte, F., Gautier, J. C., Marteau, R. and Chain, F. (1963) Consciousness disorders and akinetic mutism. Anatomo-clinical study of a bilateral paramedian softening of the cerebral peduncle and thalamus. Revue Neurologique (Paris) 109: 115–31.Google ScholarPubMed
Li, L., Miller, E. K. and Desimone, R. (1993) The representation of stimulus familiarity in anterior inferior temporal cortex. Journal of Neurophysiology 69: 1918–29.CrossRefGoogle ScholarPubMed
Li, R., Nishijo, H., Wang, Q. X.et al. (2001) Light and electron microscopic study of cholinergic and noradrenergic elements in the basolateral nucleus of the rat amygdala: Evidence for interactions between the two systems. Journal of Comparative Neurology 439: 411–25.CrossRefGoogle ScholarPubMed
Li, R., Nishijo, H., Ono, T., Ohtani, Y. and Ohtani, O. (2002) Synapses on GABAergic neurons in the basolateral nucleus of the rat amygdala: double-labeling immunoelectron microscopy. Synapse 43: 42–50.CrossRefGoogle ScholarPubMed
Li, Y. Q., Jia, H. G., Rao, Z. R. and Shi, J. W. (1990) Serotonin-, substance P- or leucine-enkephalin-containing neurons in the midbrain periaqueductal gray and nucleus raphe dorsalis send projection fibers to the central amygdaloid nucleus in the rat. Neuroscience Letters 120: 124–7.CrossRefGoogle ScholarPubMed
Liang, K. C. and McGaugh, J. L. (1983) Lesions of the stria terminals attenuate the enhancing effect of post-training epinephrine on retention of an inhibitory avoidance response. Behavioral Brain Research 9: 49–58.CrossRefGoogle Scholar
Liang, K. C., McGaugh, J. L. and Yao, H. Y. (1990) Involvement of amygdala pathways in the influence of post-training intra-amygdala norepinephrine and peripheral epinephrine on memory storage. Brain Research 508: 225–33.CrossRefGoogle ScholarPubMed
Likhtik, E., Pelletier, J. G., Paz, R. and Paré, D. (2005) Prefrontal control of the amygdala. Journal of Neuroscience 25: 7429–37.CrossRefGoogle ScholarPubMed
Lin, J. S., Sakai, K. and Jouvet, M. (1988) Evidence for histaminergic arousal mechanisms in the hypothalamus of cats. Neuropharmacology 27: 111–22.CrossRefGoogle Scholar
Lin, J. S., Sakai, K., Vanni-Mercier, G. and Jouvet, M. (1989) A critical role of the posterior hypothalamus in the mechanisms of wakefulness determined by microinjections of muscimol in freely moving cats. Brain Research 479: 225–40.CrossRefGoogle ScholarPubMed
Lin, J. S., Hou, Y., Sakai, K. and Jouvet, M. (1996) Histaminergic descending inputs to the mesopontine tegmentum and their role in the control of cortical activation and wakefulness in the cat. Journal of Neuroscience 16: 1523–37.CrossRefGoogle ScholarPubMed
Lindström, S. and Wróbel, A. (1990) Frequency dependent corticofugal excitation of principal cells in the cat's dorsal lateral geniculate nucleus. Experimental Brain Research 79: 313–8.CrossRefGoogle ScholarPubMed
Lindvall, O., Bjorklund, A. and Divac, I. (1978) Organization of catecholamine neurons projecting to the frontal cortex in the rat. Brain Research 142: 1–24.CrossRefGoogle ScholarPubMed
Linke, R., Braune, G. and Schwegler, H. (2000) Differential projection of the posterior paralaminar thalamic nuclei to the amygdaloid complex in the rat. Experimental Brain Research 134: 520–32.CrossRefGoogle ScholarPubMed
Liu, X. B. and Jones, E. G. (1999) Predominance of corticothalamic synaptic inputs to thalamic reticular nucleus neurons in the rat. Journal of Comparative Neurology 414: 67–79.3.0.CO;2-Z>CrossRefGoogle ScholarPubMed
Liu, X. B., Warren, R. A. and Jones, E. G. (1995) Synaptic distribution of afferents from reticular nucleus in ventroposterior nucleus of cat thalamus. Journal of Comparative Neurology 352: 187–202.CrossRefGoogle ScholarPubMed
Liubashina, O., Jolkkonen, E. and Pitkänen, A. (2000) Projections from the central nucleus of the amygdala to the gastric related area of the dorsal vagal complex: a Phaseolus vulgaris-leucoagglutinin study in rat. Neuroscience Letters 291: 85–8.CrossRefGoogle ScholarPubMed
Livingstone, M. S. and Hubel, D. H. (1981) Effects of sleep and arousal on the processing of visual information in the cat. Nature 291: 554–61.CrossRefGoogle ScholarPubMed
Llinás, R. R. (1988) The intrinsic electrophysiological properties of mammalian neurons: insights into central nervous system function. Science 242: 1654–64.CrossRefGoogle ScholarPubMed
Llinás, R. R. and Jahnsen, H. (1982) Electrophysiology of mammalian thalamic neurones in vitro. Nature 297: 406–8.CrossRefGoogle ScholarPubMed
Llinás, R., and Ribary, U. 1992. Rostrocaudal scan in human brain: a global characteristic for the 40 Hz response during sensory input. In Induced Rhythms in the Brain, ed. Basar, E. and Bullock, T., pp. 147–54. Boston, MA: Birkhaüser.
Llinás, R. R. and Ribary, U. (1993) Coherent 40-Hz oscillation characterizes dream state in humans. Proceedings of the National Academy of Sciences USA 90: 2078–81.CrossRefGoogle ScholarPubMed
Llinás, R., Grace, A. A. and Yarom, Y. (1991) In vitro neurons in mammalian cortical layer 4 exhibit intrinsic oscillatory activity in the 10- to 50-Hz frequency range. Proceedings of the National Academy of Sciences USA 88: 897–901.CrossRefGoogle ScholarPubMed
Llinás, R. R., Ribary, U., Jeanmonod, D., Kronberg, E. and Mitra, P. P. (1999) Thalamocortical dysrhythmia: a neurological and neuropsychiatric syndrome characterized by magnetoencephalography. Proceedings of the National Academy of Sciences USA 96: 15222–7.CrossRefGoogle ScholarPubMed
Llinás, R. R., Leznik, E. and Urbano, F. J. (2002) Temporal binding via cortical coincidence detection of specific and nonspecific thalamocortical inputs: a voltage-dependent dye-imaging study in mouse brain slices. Proceedings of the National Academy of Sciences USA 99: 449–54.CrossRefGoogle ScholarPubMed
Llinás, R. R., Urbano, F. J., Leznik, E., Ramirez, R. R. and Marle, H. H. F. (2005) Rhythmic and dysrhythmic thalamocortical dynamics: GABA systems and the edge effect. Trends in Neurosciences 28(6): 325–33.CrossRefGoogle ScholarPubMed
Logothetis, N. K. (2003) The underpinnings of the BOLD functional magnetic resonance imaging signal. Journal of Neuroscience 23: 3963–71.CrossRefGoogle ScholarPubMed
Long, M. A., Landisman, C. E. and Connors, B. W. (2004) Small clusters of electrically coupled neurons generate synchronous rhythms in the thalamic reticular nucleus. Journal of Neuroscience 24: 341–9.CrossRefGoogle ScholarPubMed
Loomis, A. L., Harvey, N. and Hobart, G. A. (1938) Distribution of disturbance patterns in the human electroencephalogram, with special reference to sleep. Journal of Neurophysiology 1: 413–30.CrossRefGoogle Scholar
Lopes da Silva, F. H., Rotterdam, A., Storm van Leeuwen, W. and Tielen, A. M. (1970) Dynamic characteristics of visual evoked potentials in the dog. II. Beta frequency selectivity in evoked potentials and background activity. Electroencephalography and Clinical Neurophysiology 29: 260–8.CrossRefGoogle ScholarPubMed
Lopes da Silva, F. H. and Storm van Leeuwen, W. (1978) The cortical alpha rhythm in dog: depth and surface profile of phase. In Architecture of the Cerebral Cortex (IBRO monograph series, vol. 3), ed. Brazier, M. A. B. and Petsche, H., pp. 319–33. New York: Raven Press.
Lopes da Silva, F. H., Lierop, T. H. M. T., Schrijer, C. F. M. and Storm Van Leeuwen, W. (1973) Organization of thalamic and cortical alpha rhythm: spectra and coherences. Electroencephalography and Clinical Neurophysiology 35: 627–39.CrossRefGoogle Scholar
Lopes da Silva, F. H., Vos, J. E., Mooibroeck, J. and Rotterdam, A. (1980) Relative contribution of intracortical and thalamo-cortical processes in the generation of alpha rhythms, revealed by partial coherence analysis. Electroencephalography and Clinical Neurophysiology 50: 449–56.CrossRefGoogle Scholar
Lopez De Armentia, M. and Sah, P. (2004) Firing properties and connectivity of neurons in the rat lateral central nucleus of the amygdala. Journal of Neurophysiology 92: 1285–94.CrossRefGoogle ScholarPubMed
Lorente de No, R. (1933) Studies on the structure of the cerebral cortex. I. The area entorhinalis. Journal für Psychologie und Neurologie 45: 381–438.Google Scholar
Loretan, K., Bissière, S. and Lüthi, A. (2004) Dopaminergic modulation of spontaneous inhibitory network activity in the lateral amygdala. Neuropharmacology 47: 631–9.CrossRefGoogle ScholarPubMed
Losier, B. J. and Semba, K. (1993) Dual projections of single cholinergic and aminergic brainstem neurons to the thalamus and basal forebrain in the rat. Brain Research 604: 41–52.CrossRefGoogle ScholarPubMed
Loughlin, S. E. and Fallon, J. H. (1983) Dopaminergic and non-dopaminergic projections to amygdala from substantia nigra and ventral tegmental area. Brain Research 262: 334–8.CrossRefGoogle ScholarPubMed
Lu, J., Greco, M. A., Shiromani, P. and Saper, C. B. (2000) Effect of lesions of the ventrolateral preoptic nucleus on NREM and REM sleep. Journal of Neuroscience 20: 3830–42.CrossRefGoogle ScholarPubMed
Lu, K. T., Walker, D. L. and Davis, M. (2001) Mitogen-activated protein kinase cascade in the basolateral nucleus of amygdala is involved in extinction of fear-potentiated startle. Journal of Neuroscience 21: NIL17–21.CrossRefGoogle ScholarPubMed
Lu, W. Y., Man, H. Y., Ju, W.et al. (2001) Activation of synaptic NMDA receptors induces membrane insertion of new AMPA receptors and LTP in cultured hippocampal neurons. Neuron 29: 243–54.CrossRefGoogle ScholarPubMed
Lucretius, T. C. (1988) On the Nature of the Universe (translated by R. E. Latham). London: Penguin Books.Google Scholar
Luebke, J. I., Greene, R. W., Semba, K.et al. (1992) Serotonin hyperpolarizes cholinergic low-threshold burst neurons in the rat laterodorsal tegmental nucleus in vitro. Proceedings of the National Academy of Sciences USA 89: 743–7.CrossRefGoogle ScholarPubMed
Lugaresi, E., Medori, R.Montagna, P.et al. (1986) Fatal familial insomnia and dysautonomia with selective degeneration of thalamic nuclei. New England Journal of Medicine 315: 997–1003.CrossRefGoogle ScholarPubMed
Luiten, P. G., Spencer, D. G., Traber, J. and Gaykema, R. P. (1985) The pattern of cortical projections from the intermediate parts of the magnocellular nucleus basalis in the rat demonstrated by tracing with Phaseolus vulgaris-leucoagglutinin. Neuroscience Letters 57: 137–42.CrossRefGoogle ScholarPubMed
Luiten, P. G., Gaykema, R. P., Traber, J. and Spencer, D. G. (1987) Cortical projection patterns of magnocellular basal nucleus subdivisions as revealed by anterogradely transported Phaseolus vulgaris leucoagglutinin. Brain Research 413: 229–50.CrossRefGoogle ScholarPubMed
Lund, J. S. (1973) Organization of neurons in the visual cortex, area 17, of the monkey (Macaca mulatta). Journal of Comparative Neurology 147: 455–96.CrossRefGoogle Scholar
Lüthi, A. and McCormick, D. A. (1998) Periodicity of thalamic synchronized oscillations: the role of Ca2+-mediated upregulation of Ih. Neuron 20: 553–63.CrossRefGoogle ScholarPubMed
Lux, H. D. and Neher, E. (1973) The equilibrium time course of [K+]o in cat cortex. Experimental Brain Research 17: 190–205.CrossRefGoogle Scholar
Lytton, W. W. and Sejnowski, T. J. (1991) Simulation of cortical pyramidal neurons synchronized by inhibitory interneurons. Journal of Neurophysiology 66: 1059–79.CrossRefGoogle ScholarPubMed
Lytton, W. W., Contreras, D., Destexhe, A. and Steriade, M. (1997) Dynamic interactions determine partial thalamic quiescence in a computer network model of spike-and-wave seizures. Journal of Neurophysiology 77: 1679–96.CrossRefGoogle Scholar
Ma, Q. P., Yin, G. F., Ai, M. K. and Han, J. S. (1991) Serotonergic projections from the nucleus raphe dorsalis to the amygdala in the rat. Neuroscience Letters 134: 21–4.CrossRefGoogle ScholarPubMed
Macchi, G., Rossi, G., Abbamondi, A. L.et al. (1997) Diffuse thalamic degeneration in fatal familial insomnia. A morphometric study. Brain Research 771: 154–8.CrossRefGoogle ScholarPubMed
Mackay, A. V. P. (1998) Letter to the Editor. Trends in Neurosciences 21: 146.CrossRefGoogle Scholar
Magill, P. J., Bolam, P. and Bevan, M. D. (2000) Relationship of activity in the subthalamic nucleus – globus pallidus network to cortical EEG. Journal of Neuroscience 20: 820–33.CrossRefGoogle Scholar
Mahanty, N. K. and Sah, P. (1998) Calcium-permeable AMPA receptors mediate long-term potentiation in interneurons in the amygdala. Nature 394: 683–7.CrossRefGoogle ScholarPubMed
Mahon, S., Deniau, J. M. and Charpier, S. (2001) Relationship between EEG potentials and intracellular activity of striatal and cortico-striatal neurons: an in vivo study under different anesthetics. Cerebral Cortex 11: 360–73.CrossRefGoogle Scholar
Mainen, Z. F. and Sejnowski, T. J. (1995) Reliability of spike timing in neocortical neurons. Science 268: 1503–6.CrossRefGoogle ScholarPubMed
Malenka, R. C. and Nicoll, R. A. (1993) NMDA-receptor-dependent synaptic plasticity: multiple forms and mechanisms. Trends in Neurosciences 16: 521–7.CrossRefGoogle ScholarPubMed
Maloney, K. J., Cape, E. G., Gotman, J. and Jones, B. E. (1997) High-frequency gamma electroencephalogram activity in association with sleep-wake states and spontaneous behaviors in the rat. Neuroscience 76: 541–55.CrossRefGoogle ScholarPubMed
Maltais, S., Coté, S., Drolet, G. and Falardeau, P. (2000) Cellular colocalization of dopamine D1 mRNA and D2 receptor in rat brain using a dopamine receptor specific polyclonal antibody. Progress in Neuro-Psychopharmacology and Biological Psychiatry 24: 1127–49.CrossRefGoogle ScholarPubMed
Manaye, K. F., Zweig, R., Wu, D.et al. (1999) Quantification of cholinergic and select non-cholinergic mesopontine neuronal populations in the human brain. Neuroscience 89: 759–70.CrossRefGoogle ScholarPubMed
Manetto, V., Medori, R., Cortelli, P.et al. (1992) Fatal familial insomnia: clinical and pathological study of five new cases. Neurology 42: 312–19.CrossRefGoogle ScholarPubMed
Manns, I. D., Mainville, L. and Jones, B. E. (2001) Evidence for glutamate, in addition to acetylcholine and GABA, neurotransmitter synthesis in basal forebrain neurons projecting to the entorhinal cortex. Neuroscience 107: 249–63.CrossRefGoogle ScholarPubMed
Mao, B. Q., Hamzei-Sichani, F., Aronov, D., Froemke, R. C. and Yuste, R. (2003) Dynamics of spontaneous activity in neocortical slices. Neuron 32: 883–98.CrossRefGoogle Scholar
Maquet, P., Péters, J. M., Aerts, J.et al. (1996) Functional neuroanatomy of human rapid-eye-movement sleep and dreaming. Nature 383: 163–6.CrossRefGoogle ScholarPubMed
Maquet, P., Degueldre, C., Delfiore, G.et al. (1997) Functional neuroanatomy of human slow wave sleep. Journal of Neuroscience 17: 2807–12.CrossRefGoogle ScholarPubMed
Maquet, P., Laureys, S., Peigneux, P.et al. (2000) Experience-dependent changes in cerebral activation during human REM sleep. Nature Neuroscience 3: 831–6.CrossRefGoogle ScholarPubMed
Maren, S. and Fanselow, M. S. (1995) Synaptic plasticity in the basolateral amygdala induced by hippocampal formation stimulation in vivo. Journal of Neuroscience 15: 7548–64.CrossRefGoogle ScholarPubMed
Maren, S. and Quirk, G. J. (2004) Neuronal signaling of fear memory. Nature Reviews Neuroscience 5: 844–52.CrossRefGoogle ScholarPubMed
Markram, H., Lübke, J., Frotscher, M., Roth, A. and Sakmann, B. (1997) Physiology and anatomy of tufted pyramidal neurons in the developing rat neocortex. Journal of Physiology (London) 500: 409–40.CrossRefGoogle ScholarPubMed
Markram, H., Wang, M. and Tsodycs, M. (1998) Differential signaling via the same axon of neocortical pyramidal neurons. Proceedings of the National Academy of Sciences USA 95: 5323–8.CrossRefGoogle ScholarPubMed
Markram, H., Toledo-Rodriguez, M., Wang, Y.et al. (2004) Interneurons of the neocortical inhibitory system. Nature Reviews Neuroscience 5: 793–807.CrossRefGoogle ScholarPubMed
Marshall, L., Mölle, M. and Born, J. (2003) Spindle and slow wave rhythms at slow wave sleep transitions are linked to strong shifts in the cortical direct current potential. Neuroscience 121: 1047–53.CrossRefGoogle ScholarPubMed
Martin, J. J. (1997) Degenerative diseases of the human thalamus. In Thalamus, vol. 2 (Experimental and Clinical Aspects), ed. Steriade, M., Jones, E. G. and McCormick, D. A., pp. 653–79. Oxford: Elsevier.Google Scholar
Martin, J. J., Yap, M., Nei, I. P. and Tan, T. E. (1983) Selective thalamic degeneration – report of a case with memory and mental disturbances. Clinical Neuropathology 2: 156–62.Google ScholarPubMed
Martina, M., Royer, S. and Paré, D. (1999) Physiological properties of central medial and central lateral amygdala neurons. Journal of Neurophysiology 82: 1843–54.CrossRefGoogle ScholarPubMed
Martina, M., Royer, S. and Paré, D. (2001) Propagation of neocortical inputs in the perirhinal cortex. Journal of Neuroscience 21: 2878–88.CrossRefGoogle ScholarPubMed
Mascagni, F., McDonald, A. J. and Coleman, J. R. (1993) Corticoamygdaloid and corticocortical projections of the rat temporal cortex: A Phaseolus vulgaris leucoagglutinin study. Neuroscience 57: 697–715.CrossRefGoogle ScholarPubMed
Massimini, M. and Amzic, F. (2001) Extracellular calcium fluctuations and intracellular potentials in the cortex during the slow sleep oscillation. Journal of Neurophysiology 85: 1346–50.CrossRefGoogle ScholarPubMed
Massimini, M., Rosanova, M. and Mariotti, M. (2003) EEG slow (∼1 Hz) waves are associated with nonstationarity of thalamo-cortical sensory processing in the sleeping human. Journal of Neurophysiology 89: 1205–13.CrossRefGoogle ScholarPubMed
Massimini, M., Huber, R., Ferrarelli, F. and Tononi, G. (2004) The sleep slow oscillation as a traveling wave. Journal of Neuroscience 24: 6862–70.CrossRefGoogle ScholarPubMed
McCarley, R. W. and Hobson, J. A. (1975) Neuronal excitability modulation over the sleep cycle: a structural and mathematical model. Science 189: 58–60.CrossRefGoogle ScholarPubMed
McCarley, R. W., Nelson, J. P. and Hobson, J. A. (1978) Ponto-geniculo-occipital (PGO) burst neurons: correlative evidence for neuronal generators of PGO waves. Science 201: 269–72.CrossRefGoogle ScholarPubMed
McCarley, R. W., Benoit, O. and Barrionuevo, G. (1983) Lateral geniculate nucleus unitary discharge in sleep and waking: state- and rate-specific aspects. Journal of Neurophysiology 50: 798–818.CrossRefGoogle ScholarPubMed
McCormick, D. A. (1992) Neurotransmitter actions in the thalamus and cerebral cortex and their role in neuromodulation of thalamocortical activity. Progress in Neurobiology 39: 337–88.CrossRefGoogle ScholarPubMed
McCormick, D. A. and Pape, H. C. (1988) Acetylcholine inhibits identified interneurons in the cat lateral geniculate nucleus. Nature 334: 246–8.CrossRefGoogle ScholarPubMed
McCormick, D. A. and Pape, H. C. (1990a) Properties of a hyperpolarization-activated cation current and its role in rhythmic oscillation in thalamic relay neurones. Journal of Physiology (London) 431: 291–318.CrossRefGoogle Scholar
McCormick, D. A. and Pape, H. C. (1990b) Noradrenergic and serotonergic modulation of a hyperpolarization-activated cation current in thalamic relay cells. Journal of Physiology (London) 431: 319–42.CrossRefGoogle Scholar
McCormick, D. A. and Wang, Z. (1991) Serotonin and noradrenaline excite GABAergic neurones of the guinea-pig and cat nucleus reticularis thalami. Journal of Physiology (London) 442: 235–55.CrossRefGoogle ScholarPubMed
McCormick, D. A. and Williamson, A. (1989) Convergence and divergence of neurotransmitter action in human cerebral cortex. Proceedings of the National Academy of Sciences USA 86: 8098–102.CrossRefGoogle ScholarPubMed
McCormick, D. A. and Williamson, A. (1991) Modulation of neuronal firing mode in cat and guinea-pig LGNd by histamine: possible cellular mechanisms of histaminergic control of arousal. Journal of Neuroscience 11: 3188–99.CrossRefGoogle ScholarPubMed
McCormick, D. A., Connors, B. W., Lighthall, J. W. and Prince, D. A. (1985) Comparative electrophysiology of pyramidal and sparsely spiny stellate neurons of the neocortex. Journal of Neurophysiology 54: 782–806.CrossRefGoogle ScholarPubMed
McDonald, A. J. (1982) Cytoarchitecture of the central amygdaloid nucleus of the rat. Journal of Comparative Neurology 208: 401–18.CrossRefGoogle ScholarPubMed
McDonald, A. J. (1985) Immunohistochemical identification of gamma-aminobutyric acid-containing neurons in the rat basolateral amygdala. Neuroscience Letters 53: 203–7.CrossRefGoogle ScholarPubMed
McDonald, A. J. (1992a) Cell types and intrinsic connections of the amygdala. In The Amygdala: Neurobiological Aspects of Emotion, Memory, and Mental Dysfunction, ed. Aggleton, J. P., pp. 67–96. New York: Wiley-Liss.Google Scholar
McDonald, A. J. (1992b) Projection neurons of the basolateral amygdala: A correlative Golgi and retrograde tract tracing study. Brain Research Bulletin 28: 179–85.CrossRefGoogle Scholar
McDonald, A. J. (1998) Cortical pathways to the mammalian amygdala. Progress in Neurobiology 55: 257–332.CrossRefGoogle ScholarPubMed
McDonald, A. J. (2003) Is there an amygdala and how far does it extend? An anatomical perspective. Annals of the New York Academy of Sciences 985: 1–21.CrossRefGoogle ScholarPubMed
McDonald, A. J. and Augustine, J. R. (1993) Localization of GABA-like immunoreactivity in the monkey amygdala. Neuroscience 52: 281–94.CrossRefGoogle ScholarPubMed
McDonald, A. J. and Betette, R. L. (2001) Parvalbumin-containing neurons in the rat basolateral amygdala: morphology and co-localization of Calbindin-D(28k). Neuroscience 102: 413–25.CrossRefGoogle Scholar
McDonald, A. J. and Mascagni, F. (1997) Projections of the lateral entorhinal cortex to the amygdala. Neuroscience 77: 445–59.CrossRefGoogle ScholarPubMed
McDonald, A. J. and Mascagni, F. (2001a) Colocalization of calcium-binding proteins and GABA in neurons of the rat basolateral amygdala. Neuroscience 105: 681–93.CrossRefGoogle Scholar
McDonald, A. J. and Mascagni, F. (2001b) Localization of the CB1 type cannabinoid receptor in the rat basolateral amygdala: high concentrations in a subpopulation of cholecystokinin-containing interneurons. Neuroscience 107: 641–52.CrossRefGoogle Scholar
McDonald, A. J. and Mascagni, F. (2002).Immunohistochemical characterization of somatostatin containing interneurons in the rat basolateral amygdala. Brain Research 943: 237–44.CrossRefGoogle ScholarPubMed
McDonald, A. J. and Mascagni, F. (2004) Parvalbumin-containing interneurons in the basolateral amygdala express high levels of the alpha1 subunit of the GABA A receptor. Journal of Comparative Neurology 473: 137–46.CrossRefGoogle Scholar
McDonald, A. J., Mascagni, F. and Guo, L. (1996) Projections of the medial and lateral prefrontal cortices to the amygdala: A Phaseolus vulgaris leucoagglutinin study in the rat. Neuroscience 71: 55–75.CrossRefGoogle ScholarPubMed
McDonald, A. J., Muller, J. F. and Mascagni, F. (2002) GABAergic innervation of alpha type II calcium/calmodulin-dependent protein kinase immunoreactive pyramidal neurons in the rat basolateral amygdala. Journal of Comparative Neurology 446: 199–218.CrossRefGoogle ScholarPubMed
McGaugh, J. L. (1966) Time-dependent processes in memory storage. Science 153: 1351–8.CrossRefGoogle ScholarPubMed
McGaugh, J. L. (2002) Memory consolidation and the amygdala: a systems perspective. Trends in Neurosciences 25: 456–61.CrossRefGoogle ScholarPubMed
McGaugh, J. L. and Gold, P. E. (1976) Modulation of memory by electrical stimulation of the brain. In Neural Mechanisms of Learning and Memory, ed. Rosenzweig, M. R. and Bennett, E. L., pp. 549–60. Cambridge, MA: MIT Press.Google Scholar
McGinty, D. J. and Harper, R. M. (1976) Dorsal raphe neurons: depression of firing during sleep in cats. Brain Research 101: 569–75.CrossRefGoogle ScholarPubMed
McKernan, M. G. and Shinnick-Gallagher, P. (1997) Fear conditioning induces a lasting potentiation of synaptic currents in vitro. Nature 390: 607–11.CrossRefGoogle ScholarPubMed
McKinney, M., Coyle, J. T. and Hedreen, J. C. (1983) Topographic analysis of the innervation of the rat neocortex and hippocampus by the basal forebrain cholinergic system. Journal of Comparative Neurology 217: 103–21.CrossRefGoogle ScholarPubMed
Mechawar, N., Cozzari, C. and Descarries, L. (2000) Cholinergic innervation in adult rat cerebral cortex: a quantitative immunocytochemical description. Journal of Comparative Neurology 428: 305–18.3.0.CO;2-Y>CrossRefGoogle ScholarPubMed
Merica, H. and Fortune, R. D. (2003) A unique pattern of sleep structure is found to be identical at all cortical sites: a neurobiological interpretation. Cerebral Cortex 13: 1044–50.CrossRefGoogle ScholarPubMed
Mesulam, M. M., Mufson, E. J., Wainer, B. H. and Levey, A. I. (1983a) Central cholinergic pathways in the rat: an overview based on an alternative nomenclature (Ch1–Ch6). Neuroscience 10: 1185–201.CrossRefGoogle Scholar
Mesulam, M.-M., Mufson, E. J., Levey, A. I. and Wainer, B. H. (1983b) Cholinergic innervation of cortex by the basal forebrain: cytochemistry and cortical connections of the septal area, diagonal band nuclei, nucleus basalis (substantia innominata), and hypothalamus in the rhesus monkey. Journal of Comparative Neurology 214: 170–97.CrossRefGoogle Scholar
Mesulam, M. M., Mufson, E. J., Levey, A. I. and Wainer, B. H. (1984) Atlas of cholinergic neurons in the forebrain and upper brainstem of the macaque based on monoclonal choline acetyltransferase immunohistochemistry and acetylcholinesterase histochemistry. Neuroscience 12: 669–86.CrossRefGoogle ScholarPubMed
Metherate, R., Cox, C. L. and Ashe, J. H. (1992) Cellular bases of neocortical activation: modulation of neural oscillations by the nucleus basalis and endogenous acetylcholine. Journal of Neuroscience 12: 4701–11.CrossRefGoogle ScholarPubMed
Meunier, M., Bachevalier, J., Mishkin, M. and Murray, E. A. (1993) Effects on visual recognition of combined and separate ablations of the entorhinal and perirhinal cortex in rhesus monkeys. Journal of Neuroscience 13: 5418–32.CrossRefGoogle ScholarPubMed
Meunier, M., Hadfield, W., Bachevalier, J. and Murray, E. A. (1996) Effects of rhinal cortex lesions combined with hippocampectomy on visual recognition memory in rhesus monkeys. Journal of Neurophysiology 75: 1190–205.CrossRefGoogle ScholarPubMed
Miettinen, N., Koivisto, E., Riekkinen, P. and Miettinen, R. (1996) Co-existence of parvalbumin and GABA in non-pyramidal neurons of the rat entorhinal cortex. Brain Research 706: 113–22.CrossRefGoogle Scholar
Miettinen, M., Pitkänen, A. and Miettinen, R. (1997) Distribution of calretinin-immunoreactivity in the rat entorhinal cortex: coexistence with GABA. Journal of Comparative Neurology 378: 363–78.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Milad, M. R. and Quirk, G. J. (2002) Neurons in medial prefrontal cortex signal memory for fear extinction. Nature 420: 70–4.CrossRefGoogle ScholarPubMed
Miller, E. K., Li, L. and Desimone, R. (1993) Activity of neurons in anterior inferior temporal cortex during a short-term memory task. Journal of Neuroscience 13: 1460–78.CrossRefGoogle ScholarPubMed
Millhouse, O. E. (1986) The intercalated cells of the amygdala. Journal of Comparative Neurology 247: 246–71.CrossRefGoogle ScholarPubMed
Minamoto, T. and Kimura, M. (2002) Participation of the thalamic CM-Pf complex in attentional orienting. Journal of Neurophysiology 87: 3090–101.CrossRefGoogle Scholar
Mineff, E. M. and Weinberg, R. J. (2000) Differential synaptic distribution of AMPA receptor subunits in the ventral posterior and reticular thalamic nuclei of the rat. Neuroscience 101: 969–82.CrossRefGoogle ScholarPubMed
Mitchell, S. and Ranck, J. B. (1980) Generation of theta rhythm in medial entorhinal cortex of freely moving rats. Brain Research 189: 49–66.CrossRefGoogle ScholarPubMed
Miyashita, Y. (1993) Inferior temporal cortex: Where visual perception meets memory. Annual Review of Neuroscience 16: 245–63.CrossRefGoogle ScholarPubMed
Mizusawa, K., Ohkoshi, N. and Sasaki, H. (1988) Degeneration of the thalamus and inferior olives associated with spongiform encephalopathy of the cerebral cortex. Clinical Neuropathology 7: 81–6.Google ScholarPubMed
Molinari, M., Minciacchi, D., Bentivoglio, M. and Macchi, G. (1985) Efferent fibers from the motor cortex terminate bilaterally in the thalamus of rats and cats. Experimental Brain Research 57: 305–12.CrossRefGoogle Scholar
Mölle, M., Marshall, L., Gais, S. and Born, J. (2002) Grouping of spindle activity during slow oscillations in human non-REM sleep. Journal of Neuroscience 22: 10941–7.CrossRefGoogle Scholar
Monckton, J. E. and McCormick, D. A. (2002) Neuromodulatory role of serotonin in the ferret thalamus. Journal of Neurophysiology 87: 2124–36.CrossRefGoogle ScholarPubMed
Monckton, J. E. and McCormick, D. A. (2003) Comparative physiological and serotoninergic properties of pulvinar neurons in the monkey, cat and ferret. Thalamus and Related Systems 2: 239–52.Google Scholar
Montero, V. (1986) Localization of gamma-aminobutyric acid (GABA) in type 3 cells and demonstration of their source to F2 terminals in the cat lateral geniculate nucleus. Journal of Comparative Neurology 254: 228–45.CrossRefGoogle ScholarPubMed
Montero, V. (1997) C-FOS induction in sensory pathways of rats exploring a novel complex environment: shifts of active thalamic reticular sectors by predominant sensory cues. Neuroscience 76: 1069–81.CrossRefGoogle ScholarPubMed
Moore, R. Y., Halaris, A. E. and Jones, B. E. (1978) Serotonin neurons of the midbrain raphe: ascending projections. Journal of Comparative Neurology 180: 417–38.CrossRefGoogle ScholarPubMed
Morgan, M. A. and LeDoux, J. E. (1995) Differential contribution of dorsal and ventral medial prefrontal cortex to the acquisition and extinction of conditioned fear in rats. Behavioral Neuroscience 109: 681–8.CrossRefGoogle ScholarPubMed
Morison, R. S. and Bassett, D. L. (1945) Electrical activity of the thalamus and basal ganglia in decorticated cats. Journal of Neurophysiology 8: 309–14.CrossRefGoogle Scholar
Morison, R. S. and Dempsey, E. W. (1942) Mechanism of thalamocortical augmentation and repetition. American Journal of Physiology 138: 297–308.Google Scholar
Morrison, J. H., and Foote, S. L. (1986) Noradrenergic and serotonergic innervation of cortical, thalamic, and tectal visual structures in old and new world monkeys. Journal of Comparative Neurology 243: 117–38.CrossRefGoogle Scholar
Morrison, J. H., Foote, S. L., O'Connor, D. and Bloom, F. E. (1982) Laminar, tangential and regional organization of the noradrenergic innervation of monkey cortex: dopamine-β-hydroxylase immunohistochemistry. Brain Research Bulletin 9: 309–19.CrossRefGoogle ScholarPubMed
Moruzzi, G. (1966) The functional significance of sleep with particular regard to the brain mechanisms underlying consciousness. In Brain and Conscious Experience, ed. Eccles, J. C., pp. 345–79. New York: Springer.Google Scholar
Moruzzi, G. (1972) The sleep-waking cycle. Ergebnisse der Physiologie 64: 1–165.Google ScholarPubMed
Moruzzi, G. and Magoun, H. W. (1949) Brain stem reticular formation and activation of the EEG. Electroencephalography and Clinical Neurophysiology 1: 455–73.CrossRefGoogle ScholarPubMed
Mountcastle, V. B. (1997) The columnar organization of the neocortex. Brain 120: 701–22.CrossRefGoogle ScholarPubMed
Mountcastle, V. B. (1998) Perceptual Neuroscience: The Cerebral Cortex. Cambridge, MA: Harvard University Press.Google Scholar
Mrzljak, L., Pappy, M., Leranth, C. and Goldman-Rakic, P. (1995) Cholinergic synaptic circuitry in the macaque prefrontal cortex. Journal of Comparative Neurology 357: 603–17.CrossRefGoogle ScholarPubMed
Mulle, C., Steriade, M. and Deschênes, M. (1985) The effects of QX-314 on thalamic neurons. Brain Research 333: 350–4.CrossRefGoogle ScholarPubMed
Muller, J., Corodimas, K. P., Fridel, Z. and LeDoux, J. E. (1997) Functional inactivation of the lateral and basal nuclei of the amygdala by muscimol infusion prevents fear conditioning to an explicit conditioned stimulus and to contextual stimuli. Behavioral Neuroscience 111: 683–91.CrossRefGoogle Scholar
Muller, J. F., Mascagni, F. and McDonald, A. J. (2003) Synaptic connections of distinct interneuronal subpopulations in the rat basolateral amygdalar nucleus. Journal of Comparative Neurology 456: 217–36.CrossRefGoogle ScholarPubMed
Mumby, D. G., Wood, E. R. and Pinel, J. (1992) Object-recognition memory is only mildly impaired in rats by lesions of the hippocampus and amygdala. Psychobiology 20: 18–27.Google Scholar
Muñoz, A., Huntsman, M. M. and Jones, E. G. (1998) GAB AB receptor gene expression in monkey thalamus. Journal of Comparative Neurology 394: 118–26.3.0.CO;2-3>CrossRefGoogle Scholar
Murakami, M., Kashiwadani, H., Kirino, Y. and Mori, K. (2005) State-dependent sensory gating in olfactory cortex. Neuron 46: 285–96.CrossRefGoogle ScholarPubMed
Murthy, V. N. and Fetz, E. E. (1992) Coherent 25- to 35-Hz oscillations in the sensorimotor cortex of awake behaving monkeys. Proceedings of the National Academy of Sciences USA 89: 5670–4.CrossRefGoogle ScholarPubMed
Musil, S. Y. and Olson, C. R. (1988) Organization of cortical and subcortical projections to medial prefrontal cortex in the cat. Journal of Comparative Neurology 272: 219–41.CrossRefGoogle ScholarPubMed
Musil, S. Y. and Olson, C. R. (1991) Cortical areas in the medial frontal lobe of the cat delineated by quantitative analysis of thalamic afferents. Journal of Comparative Neurology 308: 457–66.CrossRefGoogle ScholarPubMed
Myers, K. M. and Davis, M. (2002) Behavioral and neural analysis of extinction. Neuron 36: 567–84.CrossRefGoogle ScholarPubMed
Nadasdy, Z., Hirase, H., Czurko, A., Csicsvari, J. and Buzsáki, G. (1999) Replay and time compression of recurring spike sequences in the hippocampus. Journal of Neuroscience 19: 9497–507.CrossRefGoogle ScholarPubMed
Nader, K. (2003) Memory traces unbound. Trends in Neurosciences 26: 65–72.CrossRefGoogle ScholarPubMed
Nader, K., Majidishad, P., Amorapanth, P. and LeDoux, J. E. (2001) Damage to the lateral and central, but not other, amygdaloid nuclei prevents the acquisition of auditory fear conditioning. Learning and Memory 8: 156–63.CrossRefGoogle Scholar
Neckelmann, D., Amzica, F. and Steriade, M. (1998) Spike-wave complexes and fast components of cortically generated seizures. III. Synchronizing mechanisms. Journal of Neurophysiology 80: 1480–94.CrossRefGoogle ScholarPubMed
Nelson, J. P., McCarley, R. W. and Hobson, J. A. (1983) REM sleep burst neurons, PGO waves, and eye movement information. Journal of Neurophysiology 50: 784–97.CrossRefGoogle ScholarPubMed
Neugebauer, V. and Li, W. D. (2003) Differential sensitization of amygdala neurons to afferent inputs. Journal of Neurophysiology 89: 716–27.CrossRefGoogle ScholarPubMed
Nicoll, R. A., Malenka, R. C. and Kauer, J. A. (1990) Functional comparison of neurotransmitter receptor subtypes in mammalian central nervous system. Physiological Reviews 70: 513–65.CrossRefGoogle ScholarPubMed
Niedermeyer, E. (1993) Sleep and EEG. In Electroencephalography: Basic Principles, Clinical Applications and Related Fields (4th edition), ed. Niedermeyer, E. and Silva, F. Lopes da, pp. 153–66. Baltimore, MD: Williams & Wilkins.Google Scholar
Nielsen, T. (2000) Cognition in REM and NREM sleep. Brain and Behavioral Sciences 23: 851–66.CrossRefGoogle ScholarPubMed
Niimi, K., Matsuoka, H., Aisaka, T. and Okada, Y. (1981) Thalamic afferents to the prefrontal cortex in the cat traced with horseradish peroxidase. Journal für Hirnforschung 22: 221–41.Google ScholarPubMed
Nisenbaum, E. S., Xu, Z. C. and Wilson, C. J. (1994) Contribution of a slowly inactivating potassium current to the transition to firing of neostriatal spiny projection neurons. Journal of Neurophysiology 71: 1174–89.CrossRefGoogle ScholarPubMed
Nishimura, Y., Asahi, M., Saitoh, K.et al. (2001) Ionic mechanisms underlying burst firing of layer III sensorimotor cortical neurons of the cat: an in vitro slice study. Journal of Neurophysiology 86: 771–81.CrossRefGoogle Scholar
Nishiyama, M., Hong, K., Mikoshiba, K., Poo, M. M. and Kato, K. (2000) Calcium stores regulate the polarity and input specificity of synaptic modifications. Nature 408: 554–88.Google Scholar
Nita, D., Steriade, M. and Amzica, F. (2003) Hyperpolarisation rectification in cat lateral geniculate neurons modulated by intact corticothalamic projections. Journal of Physiology (London) 552: 325–32.CrossRefGoogle ScholarPubMed
Nitecka, L. and Ben-Ari, Y. (1987) Distribution of GABA-like immunoreactivity in the rat amygdaloid complex. Journal of Comparative Neurology 266: 45–55.CrossRefGoogle ScholarPubMed
Nitecka, L. and Frotscher, M. (1989) Organization and synaptic interconnections of GABAergic and cholinergic elements in the rat amygdaloid nuclei: single- and double-immunolabeling studies. Journal of Comparative Neurology 279: 470–88.CrossRefGoogle ScholarPubMed
Nowak, L. G., Azouz, R., Sanchez-Vives, M. V., Gray, C. M. and McCormick, D. A. (2003) Electrophysiological classes of cat primary visual cortical neurons in vivo as revealed by quantitative analyses. Journal of Neurophysiology 89: 1541–66.CrossRefGoogle ScholarPubMed
Nuñez, A., Amzica, F. and Steriade, M. (1992a) Voltage-dependent fast (20–40 Hz) oscillations in long-axoned neocortical neurons. Neuroscience 51: 7–10.CrossRefGoogle Scholar
Nuñez, A., Amzica, F. and Steriade, M. (1992b) Intrinsic and synaptically generated delta (1–4 Hz) rhythms in dorsal lateral geniculate neurons and their modulation by light-induced fast (30–70 Hz) events. Neuroscience 51: 269–84.CrossRefGoogle Scholar
Nuñez, A., Curró Dossi, R., Contreras, D. and Steriade, M. (1992c) Intracellular evidence for incompatibility between spindle and delta oscillations in thalamocortical neurons of cat. Neuroscience 48: 75–85.CrossRefGoogle Scholar
Nuñez, A., Amzica, F. and Steriade, M. (1993) Electrophysiology of cat association cortical neurons in vivo: intrinsic properties and synaptic responses. Journal of Neurophysiology 70: 418–30.CrossRefGoogle ScholarPubMed
Ojima, H. (1994) Terminal morphology and distribution of corticothalamic fibers originating from layers 5 and 6 of cat primary auditory cortex. Cerebral Cortex 4: 646–63.CrossRefGoogle ScholarPubMed
Öngür, D. and Price, J. L. (2000) The organization of networks within the orbital and medial prefrontal cortex of rats, monkeys and humans. Cerebral Cortex 10: 206–19.CrossRefGoogle ScholarPubMed
Otani, S., Daniel, H., Roisin, M. P. and Crepel, F. (2003) Dopaminergic modulation of long-term synaptic plasticity in rat prefrontal neurons. Cerebral Cortex 13: 1251–6.CrossRefGoogle ScholarPubMed
Packard, M. G. and Teather, L. A. (1998) Amygdala modulation of multiple sensory systems: hippocampus and caudate putamen. Neurobiology of Learning and Memory 69: 163–203.CrossRefGoogle Scholar
Packard, M. G., Cahill, L. and McGaugh, J. L. (1994) Amygdala modulation of hippocampal-dependent and caudate nucleus-dependent memory processes. Proceedings of the National Academy of Sciences USA 91: 8477–81.CrossRefGoogle ScholarPubMed
Pagano, R. R., Gault, M. A. and Gault, F. P. (1964) Amygdala activity: a central measure of arousal. Electroencephalography and Clinical Neurophysiology 17: 255–60.CrossRefGoogle ScholarPubMed
Palva, J. M., Palva, S. and Kaila, K. (2005) Phase synchrony among neuronal oscillations in the human cortex. Journal of Neuroscience 25: 3962–72.CrossRefGoogle ScholarPubMed
Panula, P., Pirvola, U., Auvinen, S. and Airaksinen, M. S. (1989) Histamine-immunoreactive nerve fibers in the rat brain. Neuroscience 28: 585–610.CrossRefGoogle ScholarPubMed
Panula, P., Airaksinen, M. S., Pirvola, U. and Kotilainen, E. (1990) A histamine-containing neuronal system in human brain. Neuroscience 34: 127–32.CrossRefGoogle ScholarPubMed
Pape, H. C. (1996) Queer current and pacemaker: the hyperpolarization-activated cation current in neurons. Annual Reviews of Physiology 58: 299–327.CrossRefGoogle ScholarPubMed
Pape, H. C. and Driesang, R. B. (1998) Ionic mechanisms of intrinsic oscillations in neurons of the basolateral amygdaloid complex. Journal of Neurophysiology 79: 217–26.CrossRefGoogle ScholarPubMed
Pape, H. C. and Mager, R. (1992) Nitric oxide controls oscillatory activity in thalamocortical neurons. Neuron 9: 441–8.CrossRefGoogle ScholarPubMed
Pape, H. C. and McCormick, D. A. (1995) Electrophysiological and pharmacological properties of interneurons in the cat dorsal lateral geniculate nucleus. Neuroscience 68: 1105–25.CrossRefGoogle ScholarPubMed
Pape, H. C., Budde, T., Mager, R. and Kisvárday, Z. F. (1994) Prevention of Ca2+-mediated action potentials in GABAergic local circuit neurones of rat thalamus by a transient K+ current. Journal of Physiology (London) 478: 403–22.CrossRefGoogle Scholar
Pape, H. C., Paré, D. and Driesang, R. B. (1998) Two types of intrinsic oscillations in neurons of the lateral and basolateral nuclei of the amygdala. Journal of Neurophysiology 79: 205–16.CrossRefGoogle ScholarPubMed
Paquet, M. and Smith, Y. (2000) Presynaptic NMDA receptor subunit immunoreactivity in GABAergic terminals in rat brain. Journal of Comparative Neurology 423: 330–47.3.0.CO;2-9>CrossRefGoogle ScholarPubMed
Paré, D. and Collins, D. R. (2000) Neuronal correlates of fear in the lateral amygdala: multiple extracellular recordings in conscious cats. Journal of Neuroscience 20: 2701–10.CrossRefGoogle ScholarPubMed
Paré, D. and Gaudreau, H. (1996) Projection cells and interneurons of the lateral and basolateral amygdala: distinct firing patterns and differential relation to theta and delta rhythms in conscious cats. Journal of Neuroscience 16: 3334–50.CrossRefGoogle ScholarPubMed
Paré, D. and Lang, E. J. (1998) Calcium electrogenesis in neocortical pyramidal neurons in vivo. European Journal of Neuroscience 10: 3164–70.CrossRefGoogle ScholarPubMed
Paré, D. and Llinás, R. (1995) Intracellular study of direct entorhinal inputs to field CA1 in the isolated guinea pig brain in vitro. Hippocampus 5: 115–19.CrossRefGoogle ScholarPubMed
Paré, D. and Smith, Y. (1993a) Distribution of GABA immunoreactivity in the amygdaloid complex of the cat. Neuroscience 57: 1061–76.CrossRefGoogle Scholar
Paré, D. and Smith, Y. (1993b) The intercalated cell masses project to the central and medial nuclei of the amygdala in cats. Neuroscience 57: 1077–90.CrossRefGoogle Scholar
Paré, D. and Smith, Y. (1994) GABAergic projection from the intercalated cell masses of the amygdala to the basal forebrain in cats. Journal of Comparative Neurology 344: 33–49.CrossRefGoogle ScholarPubMed
Paré, D. and Smith, Y. (1996) Thalamic collaterals of corticostriatal axons: their termination field and synaptic targets in cats. Journal of Comparative Neurology 372: 551–67.3.0.CO;2-3>CrossRefGoogle ScholarPubMed
Paré, D. and Steriade, M. (1990) Control of mamillothalamic axis by brainstem cholinergic laterodorsal tegmental afferents: possible involvement in mnemonic processes. In Brain Cholinergic Systems, ed. Steriade, M. and Biesold, D., pp. 337–54. Oxford: Oxford University Press.Google Scholar
Paré, D., Steriade, M., Deschênes, M. and Oakson, G. (1987) Physiological properties of anterior thalamic nuclei, a group devoid of inputs from the reticular thalamic nucleus. Journal of Neurophysiology 57: 1669–85.CrossRefGoogle ScholarPubMed
Paré, D., Smith, Y., Parent, A. and Steriade, M. (1988) Projections of upper brainstem cholinergic and non-cholinergic neurons of cat to intralaminar and reticular thalamic nuclei. Neuroscience 25: 69–88.CrossRefGoogle ScholarPubMed
Paré, D., Smith, Y., Parent, A. and Steriade, M. (1989) Neuronal activity of identified posterior hypothalamic neurons projecting to the brainstem peribrachial area of the cat. Neuroscience Letters 107: 145–50.CrossRefGoogle ScholarPubMed
Paré, D., Curró Dossi, R., Datta, S. and Steriade, M. (1990a) Brainstem genesis of reserpine-induced ponto-geniculo-occipital waves: an electrophysiological and morphological investigation. Experimental Brain Research 81: 533–44.CrossRefGoogle Scholar
Paré, D., Steriade, M., Deschênes, M. and Bouhassira, D. (1990b) Prolonged enhancement of anterior thalamic synaptic responsiveness by stimulation of a brainstem cholinergic group. Journal of Neuroscience 10: 20–33.CrossRefGoogle Scholar
Paré, D., Curró Dossi, R. and Steriade, M. (1991) Three types of inhibitory postsynaptic potentials generated by interneurons in the anterior thalamic complex of cat. Journal of Neurophysiology 66: 1190–204.CrossRefGoogle ScholarPubMed
Paré, D., Dong., J. and Gaudreau, H. (1995a) Amygdalo-entorhinal relations and their reflection in the hippocampal formation: generation of sharp sleep potentials. Journal of Neuroscience 15: 2482–503.CrossRefGoogle Scholar
Paré, D., Pape, H. C. and Dong, J. (1995b) Bursting and oscillating neurons of the cat basolateral amygdaloid complex in vivo: electrophysiological properties and morphological features. Journal of Neurophysiology 74: 1179–91.CrossRefGoogle Scholar
Paré, D., Smith, Y. and Paré, J. F. (1995c) Intra-amygdaloid projections of the basolateral and basomedial nuclei in the cat: Phaseolus vulgaris-leucoagglutinin anterograde tracing at the light and electron microscopic level. Neuroscience 69: 567–83.CrossRefGoogle Scholar
Paré, D., Shink, E., Gaudreau, H., Destexhe, A. and Lang, E. J. (1998a) Impact of spontaneous synaptic activity on the resting properties of cat neocortical pyramidal neurons in vivo. Journal of Neurophysiology 79: 1450–60.CrossRefGoogle Scholar
Paré, D., Lang, E. J. and Destexhe, A. (1998b) Inhibitory control of somatodendritic interactions underlying action potentials in neocortical pyramidal neurons in vivo: an intracellular and computational study. Neuroscience 84: 377–402.CrossRefGoogle Scholar
Paré, D., Collins, D. R. and Pelletier, J. G. (2002).Amygdala oscillations and the consolidation of emotional memories. Trends in Cognitive Sciences 6: 306–14.CrossRefGoogle ScholarPubMed
Paré, D., Quirk, G. J. and LeDoux, J. E. (2004) New vistas on amygdala networks in conditioned fear. Journal of Neurophysiology 92: 1–9.CrossRefGoogle ScholarPubMed
Parent, A., Descarries, L. and Beaudet, A. (1981) Organization of ascending serotonin systems in the adult rat brain. A radioautographic study after intraventricular administration of (3H)5-hydroxytryptamine. Neuroscience 6: 115–38.CrossRefGoogle Scholar
Parent, A., Paré, D., Smith, Y., and Steriade, M. (1988) Basal forebrain cholinergic and non-cholinergic projections to the thalamus and brainstem in cats and monkeys. Journal of Comparative Neurology 277: 281–301.CrossRefGoogle Scholar
Parent, M. B., Quirarte, G. L., Cahill, L. and McGaugh, J. L. (1995) Spared retention of inhibitory avoidance learning after postraining amygdala lesions. Behavioral Neuroscience 109: 803–7.CrossRefGoogle Scholar
Parmentier, R., Ohtsu, H., Djebbara-Hannas, Z.et al. (2002) Anatomical, physiological, and pharmacological characteristics of histidine decarboxylase knock-out mice: evidence for the role of brain histamine in behavioral and sleep-wake control. Journal of Neuroscience 22: 7695–711.CrossRefGoogle ScholarPubMed
Pavlides, C. and Winson, J. (1989) Influences of hippocampal place cell firing in the awake state on the activity of these cells during subsequent sleep episodes. Journal of Neuroscience 9: 2907–18.CrossRefGoogle ScholarPubMed
Pavlides, C., Greenstein, Y. J., Goudman, M. and Winson, J. (1988) Long-term potentiation in the dentate gyrus is induced preferentially on the positive phase of the theta-rhythm. Brain Research 439: 383–7.CrossRefGoogle ScholarPubMed
Pavlov, I. P. (1923) ‘Innere Hemmung’ der bedingten Reflexe und der Schlaf – ein und derselbe Prozess. Skandinavische Archive für Physiologie 44: 42–58.Google Scholar
Pavlov, I. P. (1927) Conditioned Reflexes: An Investigation of the Physiological Activity of the Cerebral Cortex. London: Oxford University Press.
Paxinos, G. and Watson, C. (1986) The Rat Brain in Stereotaxic Coordinates. New York: Academic Press.Google Scholar
Pedroarena, C. and Llinás, R. (1997) Dendritic calcium conductances generate high-frequency oscillation in thalamocortical neurons. Proceedings of the National Academy of Sciences USA 94: 724–8.CrossRefGoogle ScholarPubMed
Pelletier, J. G. and Paré, D. (2002) Uniform range of conduction times from the lateral amygdala to distributed perirhinal sites. Journal of Neurophysiology 87: 1213–21.CrossRefGoogle ScholarPubMed
Pelletier, J. G., Apergis, J. and Paré, D. (2004) Low probability transmission of neocortical and entorhinal impulses through the perirhinal cortex. Journal of Neurophysiology 91: 2079–89.CrossRefGoogle ScholarPubMed
Pelletier, J. G., Apergis-Schoute, J. and Paré, D. (2005a) Interaction between amygdala and neocortical inputs in the perirhinal cortex. Journal of Neurophysiology 94: 1837–48.CrossRefGoogle Scholar
Pelletier, J. G., Likhtik, E., Filali, M. and Paré, D. (2005b) Lasting increases in basolateral amygdala activity after emotional arousal: implications for facilitated consolidation of emotional memories. Learning and Memory 12: 96–102.CrossRefGoogle Scholar
Peña, E. and Geijo-Barrientos, E. (1996) Laminar localization, morphology, and physiological properties of pyramidal neurons that have low-threshold calcium current in the guinea-pig medial frontal cortex. Journal of Neuroscience 16: 5301–11.CrossRefGoogle ScholarPubMed
Pennartz, C. M., Uylings, H. B., Barnes, C. A. and McNaughton, B. L. (2002) Memory reactivation and consolidation during sleep: from cellular mechanisms to human performance. Progress in Brain Research 138: 143–66.CrossRefGoogle ScholarPubMed
Pennartz, C. M., Lee, E., Verheul, J.et al. (2004) The ventral striatum in off-line processing: ensemble reactivation during sleep and modulation by hippocampal ripples. Journal of Neuroscience 24: 6446–56.CrossRefGoogle ScholarPubMed
Perez-Velazquez, J. L. and Carlen, P. L. (2000) Gap junctions, synchrony and seizures. Trends in Neurosciences 23: 68–74.CrossRefGoogle ScholarPubMed
Perreault, M. C., Qin, Y., Heggelund, P. and Zhu, J. J. (2003) Postnatal development of GABAergic signalling in the rat lateral geniculate nucleus: presynaptic dendritic mechanisms. Journal of Physiology (London) 546: 137–48.CrossRefGoogle ScholarPubMed
Petersen, C. C. H., Hahn, T. T. G., Mehta, M., Grinvald, A. and Sakmann, B. (2003) Interaction of sensory responses with spontaneous depolarization in layer 2/3 barrel cortex. Proceedings of the National Academy of Sciences USA 100: 13638–43.CrossRefGoogle ScholarPubMed
Petrides, M. and Pandya, D. N. (1984) Projections to frontal cortex from posterior parietal region in rhesus monkeys. Journal of Comparative Neurology 228: 105–16.CrossRefGoogle Scholar
Petrovich, G. D. and Swanson, L. W. (1997) Projections from the lateral part of the central amygdalar nucleus to the postulated fear conditioning circuit. Brain Research 763: 247–54.CrossRefGoogle ScholarPubMed
Petsche, H., Stumpf, C. and Gogolak, G. (1962) The significance of the rabbit's septum as a relay station between the midbrain and the hippocampus. I. The control of hippocampus arousal activity by the septum cells. Electroencephalography and Clinical Neurophysiology 14: 202–11.CrossRefGoogle ScholarPubMed
Petsche, H., Pockberger, H. and Rappelsberger, P. (1984) On the search for the sources of the electroencephalogram. Neuroscience 11: 1–27.CrossRefGoogle ScholarPubMed
Peyron, C., Tighe, D. K., Pol, A. N.et al. (1998) Neurons containing hypocretin (orexin) project to multiple neuronal systems. Journal of Neuroscience 18: 9996–10015.CrossRefGoogle ScholarPubMed
Phelps, E. A. (2004) Human emotion and memory: interactions of the amygdala and hippocampal complex. Current Opinion in Neurobiology 14: 198–202.CrossRefGoogle ScholarPubMed
Phelps, E. A., Delgado, M. R., Nearing, K. I. and LeDoux, J. E. (2004) Extinction learning in humans: role of the amygdala and vmPFC. Neuron 43: 897–905.CrossRefGoogle ScholarPubMed
Pickel, V. M., Joh, T. H. and Reis, D. J. (1977) A serotonergic innervation of noradrenergic neurons in nucleus locus coeruleus: demonstration by immunocytochemical localization of the transmitter specific enzymes tyrosine and tryptophan hydroxylase. Brain Research 131: 197–214.CrossRefGoogle ScholarPubMed
Pickel, V. M., Van, B. E., Chan, J. and Cestari, D. M. (1995) Amygdala efferents form inhibitory-type synapses with a subpopulation of catecholaminergic neurons in the rat nucleus tractus solitarius. Journal of Comparative Neurology 362: 510–23.CrossRefGoogle ScholarPubMed
Pickel, V. M., Van, B. E., Chan, J. and Cestari, D. M. (1996) GABAergic neurons in rat nuclei of solitary tracts receive inhibitory-type synapses from amygdaloid efferents lacking detectable GABA-immunoreactivity. Journal of Neuroscience Research 44: 446–58.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Pikkarainen, M., Rönkkö, S., Savander, V., Insausti, R. and Pitkänen, A. (1999) Projections from the lateral, basal, and accessory basal nuclei of the amygdala to the hippocampal formation in rat. Journal of Comparative Neurology 403: 229–60.3.0.CO;2-P>CrossRefGoogle ScholarPubMed
Pinault, D., Smith, Y. and Deschênes, M. (1997) Dendrodendritic and axoaxonic synapses in the thalamic reticular nucleus of the adult rat. Journal of Neuroscience 17: 3215–33.CrossRefGoogle ScholarPubMed
Pinault, D., Leresche, N., Charpier, S.et al. (1998) Intracellular recordings in thalamic neurones during spontaneous spike and wave discharges in rats with absence epilepsy. Journal of Physiology (London) 509: 449–56.CrossRefGoogle ScholarPubMed
Pinto, A. and Sesack, S. R. (2002) Prefrontal cortex projection to the rat amygdala: ultrastructural relationship to dopamine D1 and D2 receptors. Society for Neuroscience Abstracts 28: 587.Google Scholar
Pinto, A., Fuentes, C. and Paré, D. (2006) Feedforward inhibition regulates perirhinal transmission of neocortical inputs to the entorhinal cortex: Ultrastructural study in guinea pigs. Journal of Comparative Neurology 495: 722–34.CrossRefGoogle ScholarPubMed
Pitkänen, A. (2000) Connectivity of the rat amygdaloid complex. In The Amygdala: A Functional Analysis, ed. Aggleton, J. P., pp. 31–115. Oxford: Oxford University Press.Google Scholar
Pitkänen, A. and Amaral, D. G. (1993a) Distribution of calbindin-D28k immunoreactivity in the monkey temporal lobe: the amygdaloid complex. Journal of Comparative Neurology 331: 199–224.CrossRefGoogle Scholar
Pitkänen, A. and Amaral, D. G. (1993b) Distribution of parvalbumin-immunoreactive cells and fibers in the monkey temporal lobe: the amygdaloid complex. Journal of Comparative Neurology 331: 14–36.CrossRefGoogle Scholar
Pitkänen, A. and Amaral, D. G. (1994) The distribution of GABAergic cells, fibers, and terminals in the monkey amygdaloid complex: an immunohistochemical and in situ hybridization study. Journal of Neuroscience 14: 2200–24.CrossRefGoogle ScholarPubMed
Pitkänen, A., Stefanacci, L., Farb, C. R.et al. (1995) Intrinsic connections of the rat amygdaloid complex: projections originating in the lateral nucleus. Journal of Comparative Neurology 356: 288–310.CrossRefGoogle ScholarPubMed
Pitkänen, A., Savander, V. and LeDoux, J. E. (1997) Organization of intra-amygdaloid circuitries in the rat: an emerging framework for understanding functions of the amygdala. Trends in Neurosciences 20: 517–23.CrossRefGoogle ScholarPubMed
Pitkänen, A., Pikkarainen, M., Nurminen, N. and Ylinen, A. (2000) Reciprocal connections between the amygdala and the hippocampal formation, perirhinal cortex, and postrhinal cortex in rat. Annals of the New York Academy of Sciences 911: 369–91.CrossRefGoogle ScholarPubMed
Pitkänen, A., Kelly, J. L. and Amaral, D. G. (2002) Projections from the lateral, basal, and accessory basal nuclei of the amygdala to the entorhinal cortex in the macaque monkey. Hippocampus 12: 186–205.CrossRefGoogle ScholarPubMed
Pivik, T. and Foulkes, D. (1968) NREM mentation: relation to personality, orientation time and time of night. Journal of Consultation Clinical Psychology 32: 144–51.CrossRefGoogle ScholarPubMed
Plihal, W. and Born, J. (1997) Effects of early and late nocturnal sleep on declarative and procedural memory. Journal of Cognitive Neuroscience 9: 534–47.CrossRefGoogle ScholarPubMed
Plum, F. (1991) Coma and related global disturbances of the human conscious state. In Cerebral Cortex, vol. 9 (Normal and Altered States of Function), ed. Peters, A. and Jones, E. G., pp. 359–425. New York: Plenum.CrossRefGoogle Scholar
Poggio, G. F. and Mouncastle, V. B. (1960) A study in the functional contributions of the lemniscal and spinothalamic systems to somatic sensibility: central nervous mechanisms in pain. Bulletin of the Johns Hopkins Hospital 106: 266–316.Google ScholarPubMed
Popescu, A. T., Saghyan, A. and Paré, D. (2005) Amygdala activation opens temporal windows of plasticity in the cortico-striatal pathway. Society for Neuroscience Abstracts 31: 651–4.Google Scholar
Power, A. E. (2004) Slow-wave sleep, acetylcholine, and memory consolidation. Proceedings of the National Academy of Sciences USA 101: 1795–6.CrossRefGoogle ScholarPubMed
Preuss, T. M. and Goldman-Rakic, P. S. (1987) Crossed corticothalamic and thalamocortical connections of macaque prefrontal cortex. Journal of Comparative Neurology 257: 269–81.CrossRefGoogle ScholarPubMed
Pritzel, M. and Markowitsch, H. J. (1981) Cortico-prefrontal afferents in the guinea pig. Brain Research Bulletin 7: 427–34.CrossRefGoogle ScholarPubMed
Puil, E., Meiri, H. and Yarom, Y. (1994) Resonant behavior and frequency preferences of thalamic neurons. Journal of Neurophysiology 71: 575–82.CrossRefGoogle ScholarPubMed
Purpura, D. P. (1970) Operations and processes in thalamic and synaptically related neural subsystems. In The Neuroscience: Second Study Program, ed. Schmitt, F. O., pp. 458–470. New York: Rockefeller University Press.Google Scholar
Purpura, D. P., McMurtry, J. G. and Maekawa, K. (1966) Synaptic events in ventrolateral thalamic neurons during suppression of recruiting responses by brain stem reticular stimulation. Brain Research 1: 63–76.CrossRefGoogle ScholarPubMed
Quirk, G. J., Repa, J. C. and LeDoux, J. E. (1995) Fear conditioning enhances short-latency auditory responses of lateral amygdala neurons: Parallel recordings in the freely behaving rat. Neuron 15: 1029–39.CrossRefGoogle ScholarPubMed
Quirk, G. J., Russo, G. K., Barron, J. L. and Lebron, K. (2000) The role of ventromedial prefrontal cortex in the recovery of extinguished fear. Journal of Neuroscience 20: 6225–31.CrossRefGoogle ScholarPubMed
Quirk, G. J., Likhtik, E., Pelletier, J. G. and Paré, D. (2003) Stimulation of medial prefrontal cortex decreases the responsiveness of central amygdala output neurons. Journal of Neuroscience 23: 8800–7.CrossRefGoogle ScholarPubMed
Raghavachari, S., Kahana, M. J., Rizzuto, D. S.et al. (2001) Gating of human theta oscillations by a working memory task. Journal of Neuroscience 21: 3175–83.CrossRefGoogle ScholarPubMed
Rainnie, D. G., Asprodini, E. K. and Shinnick-Gallagher, G. P. (1991) Inhibitory transmission in the basolateral amygdala. Journal of Neurophysiology 66: 999–1009.CrossRefGoogle ScholarPubMed
Rainnie, D. G., Asprodini, E. K. and Shinnick-Gallagher, G. P. (1993) Intracellular recordings from morphologically identified neurons of the basolateral amygdala. Journal of Neurophysiology 69: 1350–62.CrossRefGoogle ScholarPubMed
Ramm, P. and Smith, C. T. (1990) Rates of cerebral protein synthesis are linked to slow wave sleep in the rat. Physiology and Behavior 48: 749–53.CrossRefGoogle ScholarPubMed
Rao, Z. R., Shiosaka, S. and Tohyama, M. (1987) Origin of cholinergic fibers in the basolateral nucleus of the amygdaloid complex using sensitive double-labeling technique of retrograde biotinized tracer and immunocytochemistry. Journal für Hirnforschung 28: 553–60.Google ScholarPubMed
Rash, J. E., Staines, W. A., Yasumura, T.et al. (2000) Immunogold evidence that neuronal gap junctions in adult rat brain and spinal cord contain connexin-36 but not connexin-32 or connexin-43. Proceedings of the National Academy of Sciences USA 97: 7573–8.CrossRefGoogle ScholarPubMed
Rasler, F. E. (1984) Behavioral and electrophysiological manifestations of bombesin: excessive grooming and elimination of sleep. Brain Research 321: 187–98.CrossRefGoogle Scholar
Rasmusson, D. D., Clow, K. and Szerb, J. C. (1994) Modification of neocortical acetylcholine release and electroencephalogram desynchronization due to brainstem stimulation by drugs applied to the basal forebrain. Neuroscience 60: 665–77.CrossRefGoogle ScholarPubMed
Rasmusson, D. D., Szerb, J. C. and Jordan, J. L. (1996) Differential effects of α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid and N-methyl-D-aspartate receptor antagonists applied to the basal forebrain on cortical acetylcholine release and EEG desynchronization. Neuroscience 72: 419–27.CrossRefGoogle Scholar
Rechtschaffen, A. (1998) Current perspective on the function of sleep. Perspectives in Biology and Medicine 41: 359–90.CrossRefGoogle Scholar
Rechtschaffen, A., Verdone, P. and Wheaton, J. (1963) Reports of mental activity during sleep. Canadian Journal of Psychiatry 8: 409–14.Google ScholarPubMed
Reinagel, P., Godwin, D., Sherman, S. M. and Koch, C. (1999) Encoding of visual information by LGN bursts. Journal of Neurophysiology 81: 2558–69.CrossRefGoogle ScholarPubMed
Rempel-Clower, N. L. and Barbas, H. (1998) Topographic organization of connections between the hypothalamus and prefrontal cortex in the rhesus monkey. Journal of Comparative Neurology 398: 393–419.3.0.CO;2-V>CrossRefGoogle ScholarPubMed
Repa, J. C., Muller, J., Apergis, J.et al. (2001) Two different lateral amygdala cell populations contribute to the initiation and storage of memory. Nature Neuroscience 4: 724–31.CrossRefGoogle Scholar
Rescorla, R. A. (2004) Spontaneous recovery. Learning and Memory 11: 501–9.CrossRefGoogle ScholarPubMed
Ressler, K. J., Rothbaum, B. O., Tannenbaum, L.et al. (2004) Cognitive enhancers as adjuncts to psychotherapy: use of D-cycloserine in phobic individuals to facilitate extinction of fear. Archives of General Psychiatry 61: 1136–44.CrossRefGoogle Scholar
Ribary, U., Ioannides, A. A., Singh, K. D.et al. (1991) Magnetic field tomography of coherent thalamocortical 40-Hz oscillations in humans. Proceedings of the National Academy of Sciences USA 88: 11037–41.CrossRefGoogle ScholarPubMed
Rice, D. P. and Miller, L. S. (1995) The economic burden of affective disorders. British Journal of Psychiatry Suppl. 27: 34–42.Google Scholar
Richardson, M. P., Strange, B. A. and Dolan, R. J. (2004) Encoding of emotional memories depends on amygdala and hippocampus and their interactions. Nature Neuroscience 7: 278–85.CrossRefGoogle ScholarPubMed
Riches, I. P., Wilson, F. A. and Brown, M. W. (1991) The effects of visual stimulation and memory on neurons of the hippocampal formation and the neighboring parahippocampal gyrus and inferior temporal cortex of the primate. Journal of Neuroscience 11: 1763–79.CrossRefGoogle ScholarPubMed
Rinzel, J., Terman, D., Wang, X. J. and Ermentrout, B. (1998) Propagating activity patterns in large-scale inhibitory neuronal networks. Science 279: 1351–5.CrossRefGoogle ScholarPubMed
Robbins, T. W. and Everitt, B. J. (1995) Arousal systems and attention. In The Cognitive Neurosciences, ed. Gazzaniga, M. S., pp. 703–20. Cambridge, MA: MIT Press.
Robinson, T. E. (1980) Hippocampal rhythmic slow activity (RSA; theta): a critical analysis of selected studies and discussion of possible species-differences. Brain Research 203: 69–101.CrossRefGoogle ScholarPubMed
Roffwarg, H. P., Muzio, J. N. and Dement, W. C. (1966) Ontogenetic development of the human sleep-dream cycle. Science 152: 604–19.CrossRefGoogle ScholarPubMed
Rogan, M. T. and LeDoux, J. E. (1995) LTP is accompanied by commensurate enhancement of auditory-evoked responses in a fear conditioning circuit. Neuron 15: 127–36.CrossRefGoogle Scholar
Rogan, M. T., Stäubli, U. V. and LeDoux, J. E. (1997) Fear conditioning induces associative long-term potentiation in the amygdala. Nature 390: 604–7.CrossRefGoogle ScholarPubMed
Rolls, E. T., Miyashita, Y., Cahusac, P. M. B.et al. (1989) Hippocampal neurons in the monkey with activity related to the place in which a stimulus is shown. Journal of Neuroscience 9: 1835–45.CrossRefGoogle Scholar
Rolls, E. T., Cahusac, P., Feigenbaum, J. D. and Miyashita, Y. (1993) Responses of single neurons in the hippocampus of the macaque related to recognition memory. Experimental Brain Research 93: 299–306.CrossRefGoogle ScholarPubMed
Rolls, E. T., Inoue, K. and Browning, A. (2003) Activity of primate subgenual cingulate cortex neurons is related to sleep. Journal of Neurophysiology 90: 134–42.CrossRefGoogle ScholarPubMed
Romanski, L. M., Clugnet, M. C., Bordi, F. and LeDoux, J. E. (1993) Somatosensory and auditory convergence in the lateral nucleus of the amygdala. Behavioral Neuroscience 107: 444–50.CrossRefGoogle ScholarPubMed
Room, P. and Groenewegen, H. J. (1986a) Connections of the parahippocampal cortex. I. Cortical afferents. Journal of Comparative Neurology 251: 415–50.CrossRefGoogle Scholar
Room, P. and Groenewegen, H. J. (1986b) Connections of the parahippocampal cortex in the cat. II. Subcortical afferents. Journal of Comparative Neurology 251: 451–73.CrossRefGoogle Scholar
Room, P., Russchen, F. T., Groenewegen, H. J. and Lohman, A. H. (1985) Efferent connections of the prelimbic (area 32) and the infralimbic (area 25) cortices: an anterograde tracing study in the cat. Journal of Comparative Neurology 242: 40–55.CrossRefGoogle ScholarPubMed
Roozendaal, B. (2000) Glucocorticoids and the regulation of memory consolidation. Psychoneuroendocrinology 25: 213–38.CrossRefGoogle ScholarPubMed
Roozendaal, B. and McGaugh, J. L. (1996a) Amygdaloid nuclei lesions differentially affect glucocorticoid-induced memory enhancement in an inhibitory avoidance task. Neurobiology of Learning and Memory 65: 1–8.CrossRefGoogle Scholar
Roozendaal, B. and McGaugh, J. L. (1996b) The memory-modulatory effects of glucocorticoids depend on an intact stria terminalis. Brain Research 709: 243–50.CrossRefGoogle Scholar
Ropert, N. and Steriade, M. (1981) Input-output organization of the midbrain reticular core. Journal of Neurophysiology 46: 17–31.CrossRefGoogle ScholarPubMed
Rosanova, M. and Timofeev, I. (2005) Neuronal mechanisms mediating the variability of somatosensory evoked potentials during sleep oscillations in cats. Journal of Physiology (London) 562: 569–82.CrossRefGoogle ScholarPubMed
Rose, J. E. and Woolsey, C. N. (1948) The orbitofrontal cortex and its connections with the mediodorsal nucleus in rabbit, sheep and cat. Publications on Research of Nervous and Mental Disease 27: 210–32.Google Scholar
Rosen, J. B., Hitchcock, J. M., Sananes, C. B., Miserendino, M. and Davis, M. (1991) A direct projection from the central nucleus of the amygdala to the acoustic startle pathway: anterograde and retrograde tracing studies. Behavioral Neuroscience 105: 817–25.CrossRefGoogle ScholarPubMed
Rosene, D. L. and Va, Hoesen, G. W. (1977) Hippocampal efferents reach widespread areas of cerebral cortex and amygdala in the rhesus monkey. Science 198: 315–17.CrossRefGoogle ScholarPubMed
Rosenkranz, J. A. and Grace, A. A. (1999) Modulation of basolateral amygdala neuronal firing and afferent drive by dopamine receptor activation in vivo. Journal of Neuroscience 19: 11027–39.CrossRefGoogle ScholarPubMed
Rosenkranz, J. A. and Grace, A. A. (2001) Dopamine attenuates prefrontal cortical suppression of sensory inputs to the basolateral amygdala of rats. Journal of Neuroscience 21: 4090–103.CrossRefGoogle ScholarPubMed
Rosenkranz, J. A. and Grace, A. A. (2002) Dopamine-mediated modulation of odour-evoked amygdala potentials during Pavlovian conditioning. Nature 417: 282–7.CrossRefGoogle ScholarPubMed
Rossi, G., Macchi, G., Porro, M.et al. (1998) Fatal familial insomnia. Genetic, neuropathologic, and biochemical study of a patient from a new Italian kindred. Neurology 50: 688–92.CrossRefGoogle ScholarPubMed
Roth, M., Shaw, J. and Green, J. (1956) The form, voltage distribution and physiological significance of the K-complex. Electroencephalography and Clinical Neurophysiology 8: 385–402.CrossRefGoogle ScholarPubMed
Rothbaum, B. O. and Davis, M. (2003) Applying learning principles to the treatment of post-trauma reactions. Annals of the New York Academy of Sciences 1008: 112–21.CrossRefGoogle ScholarPubMed
Rougeul-Buser, A., Bouyer, J. J., Montaron, M. F. and Buser, P. (1983) Patterns of activities in the ventrobasal thalamus and somatic cortex SI during behavioural immobility in the awake cat: focal waking rhythms. Experimental Brain Research Suppl. 7: 69–87.CrossRefGoogle Scholar
Rowell, P. P., Volk, K. A., Li, J. and Bickford, M. E. (2003) Investigations of the cholinergic modulation of GABA release in rat thalamic slices. Neuroscience 116: 447–53.CrossRefGoogle Scholar
Roy, J. P., Clercq, M., Steriade, M. and Deschênes, M. (1984) Electrophysiology of neurons of the lateral thalamic nuclei in cat: mechanisms of long-lasting hyperpolarizations. Journal of Neurophysiology 51: 1220–35.CrossRefGoogle ScholarPubMed
Royer, S. and Paré, D. (2002) Bidirectional synaptic plasticity in intercalated amygdala neurons and the extinction of conditioned fear responses. Neuroscience 115: 455–62.CrossRefGoogle ScholarPubMed
Royer, S. and Paré, D. (2003) Conservation of total synaptic weights via inverse homo- vs. heterosynaptic LTD and LTP. Nature 422: 518–22.CrossRefGoogle Scholar
Royer, S., Martina, M. and Paré, D. (1999) An inhibitory interface gates impulse traffic between the input and output stations of the amygdala. Journal of Neuroscience 19: 10575–83.CrossRefGoogle ScholarPubMed
Royer, S., Martina, M. and Paré, D. (2000a) Bistable behavior of inhibitory neurons controlling impulse traffic through the amygdala: role of a slowly deinactivating K+ current. Journal of Neuroscience 20: 9034–9.CrossRefGoogle Scholar
Royer, S., Martina, M. and Paré, D. (2000b) Polarized synaptic interactions between intercalated neurons of the amygdala. Journal of Neurophysiology 83: 3509–18.CrossRefGoogle Scholar
Ruch-Monachon, M. A., Jalfre, M. and Haefeley, W. (1976) Drugs and PGO waves in the lateral geniculate body of the curarized rat. Archives Internationales de Pharmacodynamie et Thérapie 219: 251–346.Google Scholar
Rudolph, M. and Destexhe, A. (2003) The discharge variability of neocortical neurons during high-conductance states. Neuroscience 119: 855–73.CrossRefGoogle ScholarPubMed
Rudy, B. and McBain, C. J. (2001) Kv3 channels: voltage-gated K+ channels designed for high-frequency repetitive firing. Trends in Neurosciences 24: 517–26.CrossRefGoogle ScholarPubMed
Russchen, F. T. (1982a) Amygdalopetal projections in the cat. I. Cortical afferent connections. A study with retrograde and anterograde tracing techniques. Journal of Comparative Neurology 206: 159–79.CrossRefGoogle Scholar
Russchen, F. T. (1982b) Amygdalopetal projections in the cat. II. Subcortical afferent connections. A study with retrograde tracing techniques. Journal of Comparative Neurology 207: 157–76.CrossRefGoogle Scholar
Ruth, R. E., Collier, T. J. and Routtenberg, A. (1982) Topography between the entorhinal cortex and the dentate septotemporal axis in rats: I. Medial and intermediate entorhinal projecting cells. Journal of Comparative Neurology 209: 69–78.CrossRefGoogle ScholarPubMed
Rye, D. B., Saper, C. B., Lee, H. J. and Wainer, B. H. (1987) Pedunculopontine tegmental nucleus of the rat: cytoarchitecture, cytochemistry, and some extrapyramidal connections of the mesopontine tegmentum. Journal of Comparative Neurology 259: 483–528.CrossRefGoogle ScholarPubMed
Sacchetti, B., Lorenzini, C. A., Baldi, E., Tassoni, G. and Bucherelli, C. (1999) Auditory thalamus, dorsal hippocampus, basolateral amygdala, and perirhinal cortex role in the consolidation of conditioned freezing to context and to acoustic conditioned stimulus in the rat. Journal of Neuroscience 19: 9570–8.CrossRefGoogle ScholarPubMed
Sainsbury, R. S. (1998) Hippocampal theta: a sensory-inhibition theory of function. Neuroscience and Biobehavioral Reviews 22: 237–41.CrossRefGoogle ScholarPubMed
Sakai, K. (1985) Anatomical and physiological basis of paradoxical sleep. In Brain Mechanisms of Sleep, ed. McGinty, D. J., Morrison, A., Drucker-Colin, R. and Parmeggiani, P. L., pp. 111–37. New York: Raven.Google Scholar
Sakai, K. and Crochet, S. (2001) Differentiation of presumed serotonergic dorsal raphe neurons in relation to behavior and wake-sleep states. Neuroscience 104: 1141–55.CrossRefGoogle ScholarPubMed
Sakai, K. and Jouvet, M. (1980) Brainstem PGO-on cells projecting directly to the cat lateral geniculate nucleus. Brain Research 194: 500–5.CrossRefGoogle Scholar
Sakai, K., Salvert, D., Touret, M. and Jouvet, M. (1977a) Afferent connections of the nucleus raphe dorsalis in the cat as visualized by the horseradish peroxidase technique. Brain Research 137: 11–35.CrossRefGoogle Scholar
Sakai, K., Touret, M., Salvert, D., Leger, L. and Jouvet, M. (1977b) Afferent projections to the cat locus coeruleus as visualized by the horseradish peroxidase technique. Brain Research 119: 21–41.CrossRefGoogle Scholar
Sakai, K., El Mansari, M., Lin, J. S., Zhang, G. and Vanni-Mercier, F. (1990) The posterior hypothalamus in the regulation of wakefulness and paradoxical sleep. In The Diencephalon and Sleep, ed. Mancia, M. and Marini, G., pp. 171–98. New York: Raven.Google Scholar
Sakakura, H. (1968) Spontaneous and evoked unitary activities of cat lateral geniculate neurons in sleep and wakefulness. Japanese Journal of Physiology 18: 23–42.CrossRefGoogle ScholarPubMed
Salami, M., Itami, C., Tsumoto, T. and Kimura, F. (2003).Change in conduction velocity by regional myelination yields constant latency irrespective of distance between thalamus and cortex. Proceedings of the National Academy of Sciences USA 100: 6174–9.CrossRefGoogle ScholarPubMed
Salinas, J. A., Introini-Collison, I. B., Dalmaz, C. and McGaugh, J. L. (1997) Posttraining intraamygdala infusions of oxotremorine and propranolol modulate storage of memory for reductions in reward magnitude. Neurobiology of Learning and Memory 68: 51–9.CrossRefGoogle ScholarPubMed
Sallanon, M., Denoyer, M., Kitahama, K., Aubert, C., Gay, N. and Jouvet, M. (1989) Long-lasting insomnia induced by preoptic lesions and its transient reversal by muscimol injection into the posterior hypothalamus in the cat. Neuroscience 32: 669–83.CrossRefGoogle ScholarPubMed
Salt, T. E. (1989) Gamma-aminobutyric acid and afferent inhibition in the cat and rat ventrobasal thalamus. Neuroscience 28: 17–26.CrossRefGoogle ScholarPubMed
Samson, R. D. and Paré, D. (2005) Activity-dependent synaptic plasticity in the central nucleus of the amygdala. Journal of Neuroscience 25: 1847–55.CrossRefGoogle ScholarPubMed
Sanchez, R. and Leonard, C. S. (1996) NMDA-receptor-mediated synaptic currents in guinea pig laterodorsal tegmental neurons in vitro. Journal of Neurophysiology 76: 1101–11.CrossRefGoogle ScholarPubMed
Sanchez-Vives, M. V. and McCormick, D. A. (1997) Functional properties of perigeniculate inhibition of dorsal lateral geniculate nucleus thalamocortical neurons in vitro. Journal of Neuroscience 17: 8880–93.CrossRefGoogle ScholarPubMed
Sanchez-Vives, M. V. and McCormick, D. A. (2000) Cellular and network mechanisms of rhythmic recurrent activity in neocortex. Nature Neuroscience 3: 1027–34.CrossRefGoogle ScholarPubMed
Sanchez-Vives, M. V., Bal, T. and McCormick, D. A. (1997) Inhibitory interactions between perigeniculate GABAergic neurons. Journal of Neuroscience 17: 8894–908.CrossRefGoogle ScholarPubMed
Sanders, M. J., Wiltgen, B. J. and Fanselow, M. S. (2003) The place of the hippocampus in fear conditioning. European Journal of Pharmacology 463: 217–23.CrossRefGoogle ScholarPubMed
Santini, E., Ge, H., Ren, K., Pena de Ortiz, S. and Quirk, G. J. (2004) Consolidation of fear extinction requires protein synthesis in the medial prefrontal cortex. Journal of Neuroscience 24: 5704–10.CrossRefGoogle ScholarPubMed
Saper, C. B. (1985) Organization of cerebral cortical afferent systems in the rat. II. Hypothalamocortical projections. Journal of Comparative Neurology 237: 21–46.CrossRefGoogle ScholarPubMed
Saper, C. B. (1987) Diffuse cortical projection systems: anatomical organization and role in cortical function. In Handbook of Physiology, The Nervous System, sect. 1, vol. 5, ed. Moun, V. B. castle and Plum, F., pp. 169–210. Bethesda, MD: American Physiological Society.Google Scholar
Saper, C. B., Standaert, D. G., Currie, M. G., Schwartz, D., Geller, D. M. and Needleman, P. (1985) Atriopeptin-immunoreactive neurons in the brain: presence in cardiovascular regulatory areas. Science 227: 1047–9.CrossRefGoogle ScholarPubMed
Sarter, M. and Bruno, J. P. (2000) Cortical cholinergic inputs mediating arousal, attentional processing and dreaming: differential afferent regulation of the basal forebrain by telencephalic and brainstem afferents. Neuroscience 95: 933–52.CrossRefGoogle ScholarPubMed
Sarter, M. and Markowitsch, H. J. (1983) Convergence of basolateral amygdaloid and mediodorsal thalamic projections in different areas of the frontal cortex in the rat. Brain Research Bulletin 10: 607–22.CrossRefGoogle ScholarPubMed
Sarter, M. and Markowitsch, H. J. (1984) Collateral innervation of the medial and lateral prefrontal cortex by amygdaloid, thalamic, and brain-stem neurons. Journal of Comparative Neurology 224: 445–60.CrossRefGoogle ScholarPubMed
Satoh, K. and Fibiger, H. C. (1986) Cholinergic neurons of the laterodorsal tegmental nucleus: efferent and afferent connections. Journal of Comparative Neurology 253: 277–302.CrossRefGoogle ScholarPubMed
Saunders, M. G. and Westmoreland, B. F. (1979) The EEG in evaluation of disorders affecting the brain diffusely. In Current Practice of Clinical Electroencephalography, ed. Klass, D. W. and Daly, D. D., pp. 343–79. New York: Raven Press.Google Scholar
Saunders, R. C. and Rosene, D. L. (1988) A comparison of the efferents of the amygdala and the hippocampal formation in the rhesus monkey: I. Convergence in the entorhinal, prorhinal and perirhinal cortices. Journal of Comparative Neurology 271: 153–84.CrossRefGoogle ScholarPubMed
Saunders, R. C., Rosene, D. L. and Hoesen, G. W. (1988) Comparison of the efferents of the amygdala and the hippocampal formation in the rhesus monkey: II. Reciprocal and non-reciprocal connections. Journal of Comparative Neurology 271: 185–207.CrossRefGoogle ScholarPubMed
Savander, V., Ledoux, J. E. and Pitkänen, A. (1997) Interamygdaloid projections of the basal and accessory basal nuclei of the rat amygdaloid complex. Neuroscience 76: 725–35.CrossRefGoogle ScholarPubMed
Scalia, F. and Winans, S. S. (1975) The differential projections of the olfactory bulb and accessory olfactory bulb in mammals. Journal of Comparative Neurology 161: 31–56.CrossRefGoogle ScholarPubMed
Schafe, G. E., Nadel, N. V., Sullivan, G. M., Harris, A. and LeDoux, J. E. (1999) Memory consolidation for contextual and auditory fear conditioning is dependent on protein synthesis, PKA and MAP kinase. Learning and Memory 6: 97–110.Google ScholarPubMed
Schafe, G. E., Nader, K., Blair, H. T. and LeDoux, J. E. (2001) Memory consolidation of Pavlovian fear conditioning: a cellular and molecular perspective. Trends in Neurosciences 24: 540–6.CrossRefGoogle ScholarPubMed
Scharfman, H. E. (1991) Dentate hilar cells with dendrites in the molecular layer have lower thresholds for synaptic activation by perforant path than granule cells. Journal of Neuroscience 11: 1660–73.CrossRefGoogle ScholarPubMed
Schiess, M. C., Asprodini, E. K., Rainnie, D. G. and Shinnick-Gallagher, P. (1993) The central nucleus of the rat amygdala: in vitro intracellullar recordings. Brain Research 604: 283–97.CrossRefGoogle Scholar
Schiess, M. C., Callahan, P. M. and Zheng, H. (1999) Characterization of the electrophysiological and morphological properties of rat central amygdala neurons in vitro. Journal of Neuroscience Research 58: 663–73.3.0.CO;2-A>CrossRefGoogle ScholarPubMed
Schiff, N. D., Ribary, U., Moreno, D. R.et al. (2002) Residual cerebral activity and behavioral fragments can remain in the persistently vegetative brain. Brain 125: 1210–34.CrossRefGoogle ScholarPubMed
Schultz, W. (2002) Getting formal with dopamine and reward. Neuron 36: 241–63.CrossRefGoogle ScholarPubMed
Schwaber, J. S., Kapp, B. S., Higgins, G. A. and Rapp, P. R. (1982) Amygdaloid and basal forebrain direct connections with the nucleus of the solitary tract and the dorsal motor nucleus. Journal of Neuroscience 2: 1424–38.CrossRefGoogle ScholarPubMed
Schwindt, P. and Crill, W. (1999) Mechanisms underlying burst and regular spiking evoked by dendritic depolarization in layer 5 cortical pyramidal neurons. Journal of Neurophysiology 81: 1341–54.CrossRefGoogle ScholarPubMed
Schwindt, P. C., Spain, W. J., Foehring, R. C.et al. (1988a) Multiple potassium conductances and their functions in neurons from cat sensorimotor cortex in vitro. Journal of Neurophysiology 59: 424–49.CrossRefGoogle Scholar
Schwindt, P. C., Spain, W. J., Foehring, R. C., Chubb, M. C. and Crill, W. E. (1988b) Slow conductances in neurons from cat sensorimotor cortex in vitro and their role in slow excitability changes. Journal of Neurophysiology 59: 450–67.CrossRefGoogle Scholar
Schwindt, P. C., Spain, W. J. and Crill, W. E. (1989) Long-lasting reduction of excitability by a sodium-dependent potassium current in cat neocortical neurons. Journal of Neurophysiology 61: 233–44.CrossRefGoogle ScholarPubMed
Seamans, J. K., Nogueira, L. and Lavin, A. (2003) Synaptic basis of persistent activity in prefrontal cortex in vivo and in organotypic cultures. Cerebral Cortex 13: 1242–50.CrossRefGoogle ScholarPubMed
Segarra, J. M. (1970) Cerebral vascular disease and behaviour. I. The syndrome of the mesencephalic artery (basilary artery bifurcation). Archives of Neurology 22: 408–18.CrossRefGoogle Scholar
Seidemann, E., Meilijson, I., Abeles, M., Bergman, H. and Vaadia, E. (1996) Simultaneously recorded single units in the frontal cortex go through sequences of discrete and stable states in monkeys performing a delayed localization task. Journal of Neuroscience 16: 752–68.CrossRefGoogle ScholarPubMed
Seidenbecher, T., Laxmi, T. R., Stork, O. and Pape, H. C. (2003) Amygdalar and hippocampal theta rhythm synchronization during fear memory retrieval. Science 301: 846–50.CrossRefGoogle ScholarPubMed
Sejnowski, T. J. and Destexhe, A. (2000) Why do we sleep?Brain Research 886: 208–23.CrossRefGoogle ScholarPubMed
Semba, K. (1992) Development of central cholinergic neurons. In Handbook of Chemical Neuroanatomy, vol. 10 (Ontogeny and Transmitters of Peptides in the CNS), ed. Björkland, A., Hökfelt, T. and Tohyama, M., pp. 33–62. Amsterdam: Elsevier.Google Scholar
Semba, K. (2000) Multiple output pathways of the basal forebrain: organization, chemical heterogeneity, and role in vigilance. Behavioural Brain Research 115: 117–41.CrossRefGoogle ScholarPubMed
Semba, K. (2004) Phylogenetic and ontogenetic aspects of the basal forebrain cholinergic neurons and their innervation of the cerebral cortex. In Acetylcholine in the Cerebral Cortex (Progress in Brain Research, vol. 145), ed. Descarries, L., Krnjevic, K. and Steriade, M., pp. 3–43. Amsterdam: Elsevier.Google Scholar
Seroogy, K. B., Dangaran, K., Lim, S., Haycock, J. W. and Fallon, J. H. (1989) Ventral mesencephalic neurons containing both cholecystokinin- and tyrosine hydroxylase-like immunoreactivities project to forebrain regions. Journal of Comparative Neurology 279: 397–414.CrossRefGoogle ScholarPubMed
Sesack, S. R., Deutch, A. Y., Roth, R. H. and Bunney, B. S. (1989) Topographical organization of the efferent projections of the medial prefrontal cortex in the rat: an anterograde tract-tracing study with Phaseolus vulgaris leucoagglutinin. Journal of Comparative Neurology 290: 213–42.CrossRefGoogle ScholarPubMed
Shadlen, M. N. and Movshon, J. A. (1999) Synchrony unbound: a critical evaluation of the temporal binding hypothesis. Neuron 24: 67–77.CrossRefGoogle ScholarPubMed
Shaffery, J. P., Sinton, C. M., Bissette, G., Roffwarg, H. P. and Marks, G. A. (2002) Rapid eye movement sleep deprivation modifies expression of long-term potentiation in visual cortex of immature rats. Neuroscience 110: 431–43.CrossRefGoogle ScholarPubMed
Sheer, D. (1984) Focused arousal, 40 Hz, and dysfunction. In Selfregulation of the Brain and Behavior, ed. Ebert, T., pp. 64–84. Berlin: Springer.CrossRefGoogle Scholar
Shepherd, G. M. (2005) Perception without a thalamus: how does olfaction do it?Neuron 46: 166–8.CrossRefGoogle ScholarPubMed
Sherin, J. E., Shiromani, P. J., McCarley, R. W. and Saper, C. B. (1996) Activation of ventrolateral preoptic neurons during sleep. Science 271: 216–9.CrossRefGoogle ScholarPubMed
Sherin, J. E., Elmquist, J. K., Torrealba, F. and Saper, C. B. (1998) Innervation of histaminergic tuberomammillary neurons by GABAergic and galaninergic neurons in the ventrolateral preoptic nucleus of the rat. Journal of Neuroscience 18: 4705–21.CrossRefGoogle ScholarPubMed
Sherman, S. M. (2001) A wake-up call from the thalamus. Nature Neuroscience 4: 344–6.CrossRefGoogle ScholarPubMed
Sherrington, C. S. (1955) Man on his Nature. New York: Doubleday.Google Scholar
Shi, C. J. and Cassell, M. D. (1999) Perirhinal cortex projections to the amygdaloid complex and hippocampal formation in the rat. Journal of Comparative Neurology 406: 299–328.3.0.CO;2-9>CrossRefGoogle ScholarPubMed
Shi, S. H., Hayashi, Y., Petralia, R. S.et al. (1999) Rapid spine delivery and redistribution of AMPA receptors after synaptic NMDA receptor activation. Science 284: 1811–16.CrossRefGoogle ScholarPubMed
Shin, L. M., Orr, S. P., Carson, M. A.et al. (2004) Regional cerebral blood flow in the amygdala and medial prefrontal cortex during traumatic imagery in male and female Vietnam veterans with PTSD. Archives of General Psychiatry 61: 168–76.CrossRefGoogle ScholarPubMed
Shinonaga, Y., Takada, M. and Mizuno, N. (1994) Topographic organization of collateral projections from the basolateral amygdaloid nucleus to both the prefrontal cortex and nucleus accumbens in the rat. Neuroscience 58: 389–97.CrossRefGoogle ScholarPubMed
Shu, Y., Hasenstaub, A., Badoual, M., Bal, T. and McCormick, D. A. (2003) Barrages of synaptic activity control the gain and sensitivity of cortical neurons. Journal of Neuroscience 23: 10388–401.CrossRefGoogle ScholarPubMed
Siapas, A. G. and Wilson, M. A. (1998) Coordinated interactions between hippocampal ripples and cortical spindles during slow-wave sleep. Neuron 21: 1123–28.CrossRefGoogle ScholarPubMed
Siegel, M. and König, P. (2003) A functional gamma-band defined by stimulus-dependent synchronization in area 18 of awake behaving cats. Journal of Neuroscience 23: 4251–60.CrossRefGoogle ScholarPubMed
Sillito, A., Jones, H. E., Gerstein, G. L. and West, D. C. (1994) Feature-linked synchronization of thalamic relay cell firing induced by feedback from the visual cortex. Nature 369: 479–82.CrossRefGoogle ScholarPubMed
Silva, L. R., Gutnick, M. J. and Connors, B. W. (1991) Laminar distribution of neuronal membrane properties in neocortex of normal and reeler mouse. Journal of Neurophysiology 66: 2034–40.CrossRefGoogle ScholarPubMed
Simon, N. R., Mandshanden, I. and Lopes da Silva, F. H. (2000) A MEG study of sleep. Brain Research 860: 64–76.CrossRefGoogle Scholar
Singer, W. (1973) The effects of mesencephalic reticular stimulation on intracellular potentials of cat lateral geniculate neurons. Brain Research 61: 35–54.CrossRefGoogle Scholar
Singer, W. (1977) Control of thalamic transmission by corticofugal and ascending reticular pathways in the visual system. Physiological Reviews 57: 386–420.CrossRefGoogle ScholarPubMed
Singer, W. (1990) Role of acetylcholine in use-dependent plasticity of the visual cortex. In Brain Cholinergic Systems, ed. Steriade, M. and Biesold, D., pp. 314–36. Oxford: Oxford University Press.Google Scholar
Singer, W. (1999) Neuronal synchrony: a versatile code for the definition of relations?Neuron 24: 49–65.CrossRefGoogle ScholarPubMed
Sirota, A., Csicsvari, J., Buhl, D. and Buzsáki, G. (2003) Communication between neocortex and hippocampus during sleep in rodents. Proceedings of the National Academy of Sciences USA 100: 2065–9.CrossRefGoogle ScholarPubMed
Slotnick, S. D., Moo, L. R., Kraut, M. A., Lesser, R. P. and Hart, J. Jr. (2002) Interactions between thalamic and cortical rhythms during semantic memory recall in human. Proceedings of the National Academy of Sciences USA 99: 6440–3.CrossRefGoogle ScholarPubMed
Smith, Y. and Paré, D. (1994) Intra-amygdaloid projections of the lateral nucleus in the cat: PHA-L anterograde labeling combined with post-embedding GABA and glutamate immunocytochemistry. Journal of Comparative Neurology 342: 232–48.CrossRefGoogle Scholar
Smith, Y., Paré, D., Deschênes, M., Parent, A., and Steriade, M. (1988) Cholinergic and non-cholinergic projections from the upper brainstem core to the visual thalamus in the cat. Experimental Brain Research 70: 166–80.Google ScholarPubMed
Smith, Y., Paré, J. F. and Paré, D. (1998) Cat intraamygdaloid inhibitory network: ultrastructural organization of parvalbumin-immunoreactive elements. Journal of Comparative Neurology 391: 164–79.3.0.CO;2-0>CrossRefGoogle ScholarPubMed
Smith, Y., Paré, J. F. and Paré, D. (2000) Differential innervation of parvalbumin-immunoreactive interneurons of the basolateral amygdaloid complex by cortical and intrinsic inputs. Journal of Comparative Neurology 416: 496–508.3.0.CO;2-N>CrossRefGoogle ScholarPubMed
Snowden, R. J., Treue, S. and Andersen, R. A. (1992) The response of neurons in areas V1 and MT of the alert rhesus monkey to moving random dot patterns. Experimental Brain Research 88: 389–400.CrossRefGoogle ScholarPubMed
Sobotka, S. and Ringo, J. L. (1993) Investigation of long-term recognition and association memory in unit responses from inferotemporal cortex. Experimental Brain Research 96: 28–38.CrossRefGoogle ScholarPubMed
Soderling, T. R. and Derkach, V. A. (2000) Postsynaptic protein phosphorylation and LTP. Trends in Neurosciences 23: 75–80.CrossRefGoogle ScholarPubMed
Soltesz, I. and Crunelli, V. (1992) GABAA and pre- and post-synaptic GABAB receptor-mediated responses in the lateral geniculate nucleus. In Progress in Brain Research, vol. 90, ed. Mize, R. R., Marc, R. E. and Sillito, A. M., pp. 151–69. Amsterdam: Elsevier.Google Scholar
Soltesz, I. and Deschênes, M. (1993) Low- and high-frequency membrane potential oscillations during theta activity in CA1 and CA3 pyramidal neurons of the rat hippocampus under ketamine-xylazine anesthesia. Journal of Neurophysiology 70: 97–116.CrossRefGoogle ScholarPubMed
Soltesz, I., Lightowler, S., Leresche, N., Jassik-Gerschenfeld, D. and Crunelli, V. (1991) Two inward currents and the transformation of low-frequency oscillations of rat and cat thalamocortical cells. Journal of Physiology (London) 441: 175–97.CrossRefGoogle ScholarPubMed
Somogyi, P. (1977) A specific ‘axo-axonal’ interneuron in the visual cortex of the rat. Brain Research 136: 345–50.CrossRefGoogle ScholarPubMed
Somogyi, P. (1978) The study of Golgi stained cells and the experimental degeneration under the electron microscope: a direct method for the identification in the visual cortex of three successive links in a neuronal chain. Neuroscience 3: 167–80.CrossRefGoogle Scholar
Somogyi, P. and Cowey, A. (1981) Combined Golgi and electron microscopic study on the synapses formed by double bouquet cells in the visual cortex of the cat and monkey. Journal of Comparative Neurology 195: 547–66.CrossRefGoogle ScholarPubMed
Somogyi, P., Kisvárday, Z. F., Martin, K. A. C. and Whitteridge, D. (1983) Synaptic connections of morphologically identified and physiologically characterized large basket cells in the striate cortex of cat. Neuroscience 10: 261–94.CrossRefGoogle ScholarPubMed
Somogyi, P., Freund, T. F., Hodgson, A. J.et al. (1985) Identified axo-axonic cells are immunoreactive for GABA in the hippocampus and visual cortex of cats. Brain Research 332: 143–9.CrossRefGoogle Scholar
Somogyi, P., Tamás, G., Lujan, R. and Buhl, E. H. (1998) Salient features of synaptic organisation in the cerebral cortex. Brain Research Reviews 26: 113–35.CrossRefGoogle ScholarPubMed
Soriano, E., Martinez, A., Fariñas, I. and Frotscher, M. (1993) Chandelier cells in the hippocampal formation of the rat: the entorhinal area and subicular complex. Journal of Comparative Neurology 337: 151–67.CrossRefGoogle ScholarPubMed
Sorvari, H., Soininen, H., Paljärvi, L., Karkola, K. and Pitkänen, A. (1995).Distribution of parvalbumin-immunoreactive cells and fibers in the human amygdaloid complex. Journal of Comparative Neurology 360: 185–212.CrossRefGoogle ScholarPubMed
Soury, J. (1899) Le Système Nerveux Central. Histoire Critique des Théories et des Doctrines. Paris: Carré et Naud.Google Scholar
Spiegel, E. (1919) Die Kerne im Voderhirn des Sauger. Arbeiten der Neurologische Institut – Wien Universität 22: 418–97.Google Scholar
Spreafico, R., Amadeo, A., Angoscini, P., Panzica, F. and Battaglia, G. (1993) Branching projections from mesopontine nuclei to the nucleus reticularis and related thalamic nuclei: a double labelling study in the rat. Journal of Comparative Neurology 336: 481–92.CrossRefGoogle ScholarPubMed
Squire, L. R. and Alvarez, P. (1995) Retrograde amnesia and memory consolidation: a neurobiological perspective. Current Opinions in Neurobiology 5: 169–77.CrossRefGoogle ScholarPubMed
Standaert, D. G., Saper, C. B., Rye, D. B. and Wainer, B. H. (1986) Colocalization of atriopeptin-like immunoreactivity with choline acetyltransferase and substance P-like immunoreactivity in the pedunculopontine and laterodorsal tegmental nuclei in the rat. Brain Research 382: 163–8.CrossRefGoogle ScholarPubMed
Stanford, L. R., Friedländer, M. J. and Sherman, S. M. (1983) Morphological and physiological properties of geniculate W-cells of the cat: a comparison with X- and Y-cells. Journal of Neurophysiology 50: 582–608.CrossRefGoogle ScholarPubMed
Stefanacci, L., Suzuki, W. A. and Amaral, D. G. (1996) Organization of connections between the amygdaloid complex and the perirhinal and parahippocampal cortices in macaque monkeys. Journal of Comparative Neurology 375: 552–82.3.0.CO;2-0>CrossRefGoogle ScholarPubMed
Steininger, T. L., Gong, H., McGinty, D. and Szymusiak, R. (2001) Subregional organization of preoptic area/anterior hypothalamic projections to arousal-related monoaminergic cell groups. Journal of Comparative Neurology 429: 638–53.3.0.CO;2-Y>CrossRefGoogle ScholarPubMed
Steriade, M. (1970) Ascending control of thalamic and cortical responsiveness. International Review of Neurobiology 12: 87–144.CrossRefGoogle ScholarPubMed
Steriade, M. (1976) Cortical inhibition during sleep and waking. In Mechanisms in Transmission of Signal for Conscious Behavior, ed. Desiraju, T., pp. 209–48. Amsterdam: Elsevier.Google Scholar
Steriade, M. (1978) Cortical long-axoned cells and putative interneurons during the sleep-waking cycle. Behavioral and Brain Sciences 3: 465–514.CrossRefGoogle Scholar
Steriade, M. (1981) Mechanisms underlying cortical activation: neuronal organization and properties of the midbrain reticular core and intralaminar thalamic nuclei. In Brain Mechanisms and Perceptual Awareness, ed. Pompeiano, O. and Ajmone-Marsan, C., pp. 327–77. New York: Raven.Google Scholar
Steriade, M. (1984) The excitatory-inhibitory response sequence of thalamic and neocortical cells: state-related changes and regulatory systems. In Dynamic Aspects of Neocortical Function, ed. Edelman, G. M., Gall, W. E. and Cowan, W. M., pp. 107–57. New York: Wiley.Google Scholar
Steriade, M. (1991) Alertness, quiet sleep, dreaming. In Cerebral Cortex, vol. 9 (Normal and Altered States of Function), ed. Peters, A. and Jones, E. G., pp. 279–357. New York: Plenum.CrossRefGoogle Scholar
Steriade, M. (1995a) Two channels in the cerebellothalamocortical system. Journal of Comparative Neurology 354: 57–70.CrossRefGoogle Scholar
Steriade, M. (1995b) Brain activation, then (1949) and now: coherent fast rhythms in corticothalamic networks. Archives Italiennes de Biologie 134: 5–20.Google Scholar
Steriade, M. (1997a) Synchronized activities of coupled oscillators in the cerebral cortex and thalamus at different levels of vigilance. Cerebral Cortex 7: 583–604.CrossRefGoogle Scholar
Steriade, M. (1997b) Thalamic substrates of disturbances in states of vigilance and consciousness in humans. In Thalamus, vol. 2 (Experimental and Clinical Aspects), ed. Steriade, M., Jones, E. G. and McCormick, D. A., pp. 721–42. Oxford: Elsevier.Google Scholar
Steriade, M. (1999) Coherent oscillations and short-term plasticity in corticothalamic networks. Trends in Neurosciences 22: 337–45.CrossRefGoogle ScholarPubMed
Steriade, M. (2000) Corticothalamic resonance, states of vigilance, and mentation. Neuroscience 101: 243–76.CrossRefGoogle ScholarPubMed
Steriade, M. (2001a) The Intact and Sliced Brain. Cambridge, MA: MIT Press.Google Scholar
Steriade, M. (2001b) Impact of network activities on neuronal properties in corticothalamic systems. Journal of Neurophysiology 86: 1–39.CrossRefGoogle Scholar
Steriade, M. (2001c) The GABAergic reticular nucleus: a preferential target of corticothalamic projections. Proceedings of the National Academy of Sciences USA 98: 3625–7.CrossRefGoogle Scholar
Steriade, M. (2001d) To burst, or rather, not to burst. Nature Neuroscience 4: 671.CrossRefGoogle Scholar
Steriade, M. (2003a) Neuronal Substrates of Sleep and Epilepsy. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Steriade, M. (2003b) The corticothalamic system in sleep. Frontiers in Bioscience 8: 878–99.CrossRefGoogle Scholar
Steriade, M. (2003c) Presynaptic dendrites of thalamic local-circuit neurons and sculpting inhibition during activated states. Journal of Physiology (London) 546: 1.CrossRefGoogle Scholar
Steriade, M. (2004a) Neocortical neuronal classes are flexible entities. Nature Reviews Neuroscience 5: 121–34.CrossRefGoogle Scholar
Steriade, M. (2004b).Local gating of information processing through the thalamus. Neuron 41: 493–4.CrossRefGoogle Scholar
Steriade, M. (2004c).Slow-wave sleep: serotonin, neuronal plasticity, and seizures. Archives Italiennes de Biologie 142: 359–67.Google Scholar
Steriade, M. (2005a) Cellular substrates of brain rhythms. In Electroencephalography: Basic Principles, Clinical Applications, and Related Fields (5th edition), ed. Niedermeyer, E. and Silva, F. Lopes Da, pp. 31–83. Baltimore, MD: Williams & Wilkins.Google Scholar
Steriade, M. (2005b) Neuronal substrates of spike-wave seizures and hypsarrhythmia in corticothalamic systems. Advances in Neurology 27: 149–54.Google Scholar
Steriade, M. (2006) Brainstem-thalamic neurons implicated in hallucinations. Behavioral and Brain Sciences, in press.Google Scholar
Steriade, M. and Amzic, F. (1996) Intracortical and corticothalamic coherency of fast spontaneous oscillations. Proceedings of the National Academy of Sciences USA 93: 2533–8.CrossRefGoogle ScholarPubMed
Steriade, M. and Amzica, F. (1998) Coalescence of sleep rhythms and their chronology in corticothalamic networks. Sleep Research Online 1: 1–10.Google ScholarPubMed
Steriade, M. and Buzsáki, G. (1990). Parallel activation of thalamic and cortical neurons by brainstem and basal forebrain cholinergic systems. In Brain Cholinergic Systems, ed. Steriade, M. and Biesold, D., pp. 3–63. Oxford: Oxford University Press.Google Scholar
Steriade, M. and Contreras, D. (1993) Sleep oscillations in interacting thalamocortical networks. In Thalamic Networks for Relay and Modulation, ed. Minciacchi, D., Molinari, M., Macchi, G. and Jones, E. G., pp. 385–94. Oxford: Pergamon Press.Google Scholar
Steriade, M. and Contreras, D. (1995) Relations between cortical and thalamic cellular events during transition from sleep pattern to paroxysmal activity. Journal of Neuroscience 15: 623–42.CrossRefGoogle Scholar
Steriade, M. and Demetrescu, M. (1960) Unspecific systems of inhibition and facilitation of potentials evoked by intermittent light. Journal of Neurophysiology 23: 602–17.CrossRefGoogle Scholar
Steriade, M. and Demetrescu, M. (1967) Simulation of peripheral sensory input by electrical pulse trains applied to specific afferent pathways. Experimental Neurology 19: 265–77.CrossRefGoogle ScholarPubMed
Steriade, M. and Deschênes, M. (1974) Inhibitory processes and interneuronal apparatus in motor cortex during sleep and waking. II. Recurrent and afferent inhibition of pyramidal tract neurons. Journal of Neurophysiology 37: 1093–113.CrossRefGoogle ScholarPubMed
Steriade, M. and Deschênes, M. (1984) The thalamus as a neuronal oscillator. Brain Research Reviews 8: 1–63.CrossRefGoogle Scholar
Steriade, M. and Deschênes, M. (1987) Inhibitory processes in the thalamus. Journal of Mind and Behavior 8: 559–72.Google Scholar
Steriade, M. and Deschênes, M. (1988) Intrathalamic and brainstem-thalamic networks involved in resting and alert states. In Cellular Thalamic Mechanisms, ed. Bentivoglio, M., Macchi, G. and Spreafico, R., pp. 37–62. Amsterdam: Elsevier.Google Scholar
Steriade, M. and Glenn, L. L. (1982) Neocortical and caudate projections of intralaminar thalamic neurons and their synaptic excitation from the midbrain reticular core. Journal of Neurophysiology 48: 352–71.CrossRefGoogle ScholarPubMed
Steriade, M. and Hobson, J. A. (1976) Neuronal activity during the sleep-waking cycle. Progress in Neurobiology 6: 155–376.CrossRefGoogle ScholarPubMed
Steriade, M. and Llinás, R. R. (1988) The functional states of the thalamus and the associated neuronal interplay. Physiological Reviews 68: 649–742.CrossRefGoogle ScholarPubMed
Steriade, M. and McCarley, R. W. (2005) Brain Control of Wakefulness and Sleep (2nd edition). New York: Springer-Kluwer.Google Scholar
Steriade, M. and Morin, D. (1981) Reticular influences on primary and augmenting responses in the somatosensory cortex. Brain Research 205: 67–80.CrossRefGoogle ScholarPubMed
Steriade, M. and Timofeev, I. (1997) Short-term plasticity during intrathalamic augmenting responses in decorticated cats. Journal of Neuroscience 17: 3778–95.CrossRefGoogle ScholarPubMed
Steriade, M. and Timofeev, I. (2001) Corticothalamic operations through prevalent inhibition of thalamocortical neurons. Thalamus and Related Systems 1: 225–36.Google Scholar
Steriade, M. and Timofeev, I. (2003) Neuronal plasticity in thalamocortical networks during sleep and waking oscillations. Neuron 37: 563–76.CrossRefGoogle ScholarPubMed
Steriade, M., Belekhova, M. and Apostol, V. (1968) Reticular potentiation of cortical flash-evoked afterdischarge. Brain Research 11: 276–80.CrossRefGoogle ScholarPubMed
Steriade, M., Iosif, G. and Apostol, V. (1969) Responsiveness of thalamic and cortical motor relays during arousal and various stages of sleep. Journal of Neurophysiology 32: 251–65.CrossRefGoogle Scholar
Steriade, M., Apostol, V. and Oakson, G. (1971) Control of unitary activities in cerebellothalamic pathway during wakefulness and synchronized sleep. Journal of Neurophysiology 34: 389–413.CrossRefGoogle ScholarPubMed
Steriade, M., Wyzinski, P. and Apostol, V. (1972) Corticofugal projections governing rhythmic thalamic activity. In Corticothalamic Projections and Sensorimotor Activities, ed. Frigyesi, T. L., Rinvik, E. and Yahr, M. D., pp. 221–72. New York: Raven Press.Google Scholar
Steriade, M., Deschênes, M. and Oakson, G. (1974a) Inhibitory processes and interneuronal apparatus in motor cortex during sleep and waking. I. Background firing and synaptic responsiveness of pyramidal tract neurons and interneurons. Journal of Neurophysiology 37: 1065–92.CrossRefGoogle Scholar
Steriade, M., Deschênes, M., Wyzinski, P. and Hallé, J. P. (1974b) Input-output organization of the motor cortex during sleep and waking. In Basic Sleep Mechanisms, ed. Petre-Quadens, O. and Schlag, J., pp. 144–200. New York: Academic Press.Google Scholar
Steriade, M., Oakson, G. and Diallo, A. (1977) Reticular influences on lateralis posterior thalamic neurons. Brain Research 131: 55–71.CrossRefGoogle ScholarPubMed
Steriade, M., Kitsikis, A. and Oakson, G. (1979) Excitatory-inhibitory processes in parietal association neurons during reticular activation and sleep-waking cycle. Sleep 1: 339–55.CrossRefGoogle ScholarPubMed
Steriade, M., Oakson, G. and Ropert, N. (1982a) Firing rates and patterns of midbrain reticular neurons during steady and transitional states of the sleep-waking cycle. Experimental Brain Research 46: 37–51.CrossRefGoogle Scholar
Steriade, M., Parent, A., Ropert, N. and Kitsikis, A. (1982b) Zona incerta and lateral hypothalamic afferents to the midbrain reticular core of the cat – an HRP and electrophysiological study. Brain Research 238: 13–28.CrossRefGoogle Scholar
Steriade, M., Parent, A. and Hada, J. (1984a) Thalamic projections of nucleus reticularis thalami of cat: a study using retrograde transport of horseradish peroxidase and double fluorescent tracers. Journal of Comparative Neurology 229: 531–47.CrossRefGoogle Scholar
Steriade, M., Sakai, K. and Jouvet, M. (1984b) Bulbothalamic neurons related to thalamocortical activation processes during paradoxical sleep. Experimental Brain Research 54: 463–75.CrossRefGoogle Scholar
Steriade, M., Deschênes, M., Domich, L. and Mulle, C. (1985) Abolition of spindle oscillations in thalamic neurons disconnected from nucleus reticularis thalami. Journal of Neurophysiology 54: 1473–97.CrossRefGoogle ScholarPubMed
Steriade, M., Domich, L. and Oakson, G. (1986) Reticularis thalami neurons revisited: activity changes during shifts in states of vigilance. Journal of Neuroscience 6: 68–81.CrossRefGoogle ScholarPubMed
Steriade, M., Domich, L., Oakson, G. and Deschênes, M. (1987a) The deafferented reticular thalamic nucleus generates spindle rhythmicity. Journal of Neurophysiology 57: 260–73.CrossRefGoogle Scholar
Steriade, M., Parent, A., Paré, D. and Smith, Y. (1987b) Cholinergic and non-cholinergic neurons of cat basal forebrain project to reticular and mediodorsal thalamic nuclei. Brain Research 408: 372–76.CrossRefGoogle Scholar
Steriade, M., Paré, D., Parent, A. and Smith, Y. (1988) Projections of cholinergic and non-cholinergic neurons of the brainstem core to relay and associational thalamic nuclei in the cat and macaque monkey. Neuroscience 25: 47–67.CrossRefGoogle ScholarPubMed
Steriade, M., Paré, D., Bouhassira, D., Deschênes, M. and Oakson, G. (1989) Phasic activation of lateral geniculate and perigeniculate neurons during sleep with ponto-geniculo-occipital spikes. Journal of Neuroscience 9: 2215–29.CrossRefGoogle Scholar
Steriade, M., Datta, S., Paré, D., Oakson, G. and Curró Dossi, R. (1990a) Neuronal activities in brainstem cholinergic nuclei related to tonic activation processes in thalamocortical systems. Journal of Neuroscience 10: 2541–59.CrossRefGoogle Scholar
Steriade, M., Jones, E. G. and Llinás, R. R. (1990b) Thalamic Oscillations and Signaling. New York: Wiley-Interscience.Google Scholar
Steriade, M., Gloor, P., Llinás, R. R., Lopes da Silva, F. H. and Mesulam, M. M. (1990c) Basic mechanisms of cerebral rhythmic activities. Electroencephalography and Clinical Neurophysiology 76: 481–508.CrossRefGoogle Scholar
Steriade, M., Paré, D., Datta, S., Oakson, G. and Curró Dossi, R. (1990d) Different cellular types in mesopontine cholinergic nuclei related to ponto-geniculo-occipital waves. Journal of Neuroscience 10: 2560–79.CrossRefGoogle Scholar
Steriade, M., Curró Dossi, R. and Nuñez, A. (1991a) Network modulation of a slow intrinsic oscillation of cat thalamocortical neurons implicated in sleep delta waves: cortical potentiation and brainstem cholinergic suppression. Journal of Neuroscience 11: 3200–17.CrossRefGoogle Scholar
Steriade, M., Curró Dossi, R., Paré, D. and Oakson, G. (1991b) Fast oscillations (20–40 Hz) in thalamocortical systems and their potentiation by mesopontine cholinergic nuclei in the cat. Proceedings of the National Academy of Sciences USA 88: 4396–400.CrossRefGoogle Scholar
Steriade, M., Amzica, F. and Nuñez, A. (1993a) Cholinergic and noradrenergic modulation of the slow (∼0.3 Hz) oscillation in neocortical cells. Journal of Neurophysiology 70: 1384–400.CrossRefGoogle Scholar
Steriade, M., Contreras, D., Curró Dossi, R. and Nuñez, A. (1993b) The slow (<1 Hz) oscillation in reticular thalamic and thalamocortical neurons: scenario of sleep rhythm generation in interacting thalamic and neocortical networks. Journal of Neuroscience 13: 3284–99.CrossRefGoogle Scholar
Steriade, M., Curró Dossi, R. and Contreras, D. (1993c) Electrophysiological properties of intralaminar thalamocortical cells discharging rhythmic (∼40 Hz) spike-bursts at ∼1000 Hz during waking and rapid eye movement sleep. Neuroscience 56: 1–9.CrossRefGoogle Scholar
Steriade, M., McCormick, D. A. and Sejnowski, T. J. (1993d) Thalamocortical oscillation in the sleeping and aroused brain. Science 262: 679–85.CrossRefGoogle Scholar
Steriade, M., Nuñez, A. and Amzica, F. (1993e) A novel slow (<1 Hz) oscillation of neocortical neurons in vivo: depolarizing and hyperpolarizing components. Journal of Neuroscience 13: 3252–65.CrossRefGoogle Scholar
Steriade, M., Nuñez, A. and Amzica, F. (1993f) Intracellular analysis of relations between the slow (<1 Hz) neocortical oscillation and other sleep rhythms. Journal of Neuroscience 13: 3266–83.CrossRefGoogle Scholar
Steriade, M., Amzica, F. and Contreras, D. (1994a) Cortical and thalamic cellular correlates of electroencephalographic burst-suppression. Electroencephalography and Clinical Neurophysiology 90: 1–16.CrossRefGoogle Scholar
Steriade, M., Contreras, D. and Amzica, F. (1994b) Synchronized sleep oscillations and their paroxysmal developments. Trends in Neuroscience 17: 199–208.CrossRefGoogle Scholar
Steriade, M., Amzica, F. and Contreras, D. (1996a) Synchronization of fast (30–40 Hz) spontaneous cortical rhythms during brain activation. Journal of Neuroscience 16: 392–417.CrossRefGoogle Scholar
Steriade, M., Contreras, D., Amzica, F. and Timofeev, I. (1996b) Synchronization of fast (30–40 Hz) spontaneous oscillations in intrathalamic and thalamocortical networks. Journal of Neuroscience 16: 2788–808.CrossRefGoogle Scholar
Steriade, M., Jones, E. G. and McCormick, D. A. (1997) Thalamus, vol. 1 (Organisation and Function). Oxford: Elsevier.Google Scholar
Steriade, M., Timofeev, I., Dürmüller, N. and Grenier, F. (1998a) Dynamic properties of corticothalamic neurons and local cortical interneurons generating fast rhythmic (30–40 Hz) spike bursts. Journal of Neurophysiology 79: 483–90.CrossRefGoogle Scholar
Steriade, M., Timofeev, I., Grenier, F. and Dürmüller, N. (1998b) Role of thalamic and cortical neurons in augmenting responses: dual intracellular recordings in vivo. Journal of Neuroscience 18: 6425–43.CrossRefGoogle Scholar
Steriade, M., Timofeev, I. and Grenier, F. (2001a) Natural waking and sleep states: a view from inside neocortical neurons. Journal of Neurophysiology 85: 1969–85.CrossRefGoogle Scholar
Steriade, M., Timofeev, I. and Grenier, F. (2001b) Intrinsic, antidromic and synaptic excitability of cortical neurons during natural waking-sleep cycle. Society for Neuroscience Abstracts 27: 240.Google Scholar
Stickgold, R., James, L. and Hobson, J. A. (2000a) Visual discrimination learning requires sleep after training. Nature Neuroscience 3: 1237–8.CrossRefGoogle Scholar
Stickgold, R., Whitbee, D., Schirmer, B., Patel, V. and Hobson, J. A. (2000b) Visual discrimination improvement. A multi-step process occurring during sleep. Journal of Cognitive Neuroscience 12: 246–54.CrossRefGoogle Scholar
Stone, V. E., Baron-Cohen, S. and Knight, R. T. (1998) Frontal lobe contributions in theory of the mind. Journal of Cognitive Neuroscience 10: 640–56.CrossRefGoogle ScholarPubMed
Storm, J. F. (1988) Temporal integration by a slowly inactivating K+ current in hippocampal neurons. Nature 336: 379–81.CrossRefGoogle ScholarPubMed
Stratford, K. J., Tarczy-Hornoch, K., Martin, K. A. C., Bannister, N. J. and Jack, J. J. B. (1996) Excitatory synaptic inputs to spiny stellate cells in cat visual cortex. Nature 382: 258–60.CrossRefGoogle ScholarPubMed
Stuart, G. and Sakmann, B. (1994) Active propagation of somatic action potentials into neocortical pyramidal cell dendrites. Nature 367: 69–72.CrossRefGoogle ScholarPubMed
Stuart, G., Spruston, N., Sakmann, B. and Häusser, M. (1997) Action potential initiation and backpropagation in neurons of the mammalian CNS. Trends in Neurosciences 20: 125–31.CrossRefGoogle ScholarPubMed
Sugihara, I., Lang, E. J. and Llinás, R. (1993) Uniform olivocerebellar conduction time underlies Purkinje cell complex spike synchronicity in the rat cerebellum. Journal of Physiology (London) 470: 243–71.CrossRefGoogle ScholarPubMed
Sun, N. and Cassell, M. D. (1993) Intrinsic GABAergic neurons in the rat central extended amygdala. Journal of Comparative Neurology 330: 381–404.CrossRefGoogle ScholarPubMed
Sutton, R. E., Koob, G. F., Moal, M., Rivier, J. and Vale, W. (1982) Corticotropin releasing factor produces behavioral activation in rats. Nature 297: 31–3.CrossRefGoogle ScholarPubMed
Suzuki, W. A. (1996) The anatomy, physiology and functions of the perirhinal cortex. Current Opinion in Neurobiology 6: 179–86.CrossRefGoogle ScholarPubMed
Suzuki, W. A. and Porteros, A. (2002) Distribution of calbindin D-28k in the entorhinal, perirhinal, and parahippocampal cortices of the macaque monkey. Journal of Comparative Neurology 451: 392–412.CrossRefGoogle ScholarPubMed
Suzuki, W. A., Zola-Morgan, S., Squire, L. R. and Amaral, D. G. (1993) Lesions of the perirhinal and parahippocampal cortices in the monkey produce long-lasting memory impairment in the visual and tactual modalities. Journal of Neuroscience 13: 2430–51.CrossRefGoogle ScholarPubMed
Svoboda, K., Denk, W., Kleinfeld, D. and Tank, D. W. (1997) In vivo dendritic calcium dynamics in neocortical neurons. Nature 385: 161–5.CrossRefGoogle Scholar
Svoboda, K., Helmchen, F., Denk, W. and Tank, D. W. (1999) Spread of dendritic excitation in layer 2/3 pyramidal neurons in rat barrel cortex in vivo. Nature Neuroscience 2: 65–73.CrossRefGoogle ScholarPubMed
Swadlow, H. A. and Gusev, A. G. (2001) The impact of ‘bursting’ thalamic impulses at a neocortical synapse. Nature Neuroscience 4: 402–8.CrossRefGoogle Scholar
Swanson, L. W. (1992) Brain Maps: Structure of the Rat Brain. Amsterdam: Elsevier.Google Scholar
Swanson, L. W. and Hartman, B. K. (1975) The central adrenergic system. An immunofluorescence study of the location of cell bodies and their efferent connections in the rat utilizing dopamine-beta-hydroxylase as a marker. Journal of Comparative Neurology 163: 467–505.CrossRefGoogle ScholarPubMed
Swanson, L. W. and Petrovich, G. D. (1998) What is the amygdala?Trends in Neurosciences 21: 323–31.CrossRefGoogle ScholarPubMed
Swanson, L. W., Mogenson, G. J., Simerly, R. B. and Wu, M. (1987) Anatomical and electrophysiological evidence for a projection from the medial preoptic area to the ‘mesencephalic and subthalamic locomotor regions’ in the rat. Brain Research 405: 108–22.CrossRefGoogle ScholarPubMed
Szymusiak, R. and McGinty, D. (1986) Sleep-related neuronal discharge in the basal forebrain of cats. Brain Research 370: 82–92.CrossRefGoogle ScholarPubMed
Szymusiak, R., Steininger, T., Alam, N. and McGinty, D. (2001) Preoptic area sleep-regulating mechanisms. Archives Italiennes de Biologie 139: 77–92.Google ScholarPubMed
Takada, M. (1990) The A11 catecholamine cell group: another origin of the dopaminergic innervation of the amygdala. Neuroscience Letters 118: 132–5.CrossRefGoogle ScholarPubMed
Takagishi, M. and Chiba, T. (1991) Efferent projections of the infralimbic (area 25) region of the medial prefrontal cortex in the rat: an anterograde tracer PHA-L study. Brain Research 566: 26–36.CrossRefGoogle ScholarPubMed
Takakusaki, K. and Kitai, S. T. (1997) Ionic mechanisms involved in the spontaneous firing of tegmental pedunculopontine nucleus neurons of the rat. Neuroscience 78: 771–94.CrossRefGoogle ScholarPubMed
Takakusaki, K., Shiroyama, T. and Kitai, S. T. (1997) Two types of cholinergic neurons in he rat tegmental pedunculopontine nucleus: electrophysiological and morphological characterization. Neuroscience 79: 1089–99.CrossRefGoogle Scholar
Takeuchi, Y., McLean, J. H. and Hopkins, D. A. (1982) Reciprocal connections between the amygdala and parabrachial nuclei: ultrastructural demonstration by degeneration and axonal transport of horse radish peroxidase in the cat. Brain Research 239: 583–8.CrossRefGoogle Scholar
Takeuchi, Y., Matsushima, S., Matsuchima, R. and Hopkins, D. A. (1983) Direct amygdaloid projections to the dorsal motor nucleus of the vagus nerve: a light and electron microscopic study in the rat. Brain Research 280: 143–7.CrossRefGoogle ScholarPubMed
Takita, M., Izaki, Y., Jay, T. M., Kaneko, H. and Suzuki, S. S. (1999) Induction of stable long-term depression in vivo in the hippocampal-prefrontal cortex pathway. European Journal of Neuroscience 11: 4145–8.CrossRefGoogle ScholarPubMed
Tamás, G., Buhl, E. H. and Somogyi, P. (1997) Massive autaptic self-innervation of GABAergic neurons in cat visual cortex. Journal of Neuroscience 17: 6352–64.CrossRefGoogle ScholarPubMed
Tamás, G., Somogyi, P. and Buhl, E. H. (1998) Differentially interconnected networks of GABAergic interneurons in the visual cortex of the cat. Journal of Neuroscience 18: 4255–70.CrossRefGoogle ScholarPubMed
Tanabe, T., Yarita, H., Iino, M., Ooshima, Y. and Takagi, S. F. (1975a) An olfactory projection area in orbitofrontal cortex of the monkey. Journal of Neurophysiology 38: 1269–83.CrossRefGoogle Scholar
Tanabe, T., Iino, M. and Takagi, S. F. (1975b) Discrimination of odors in olfactory bulb, pyriform-amygdaloid areas, and orbitofrontal cortex of the monkey. Journal of Neurophysiology 38: 1284–96.CrossRefGoogle Scholar
Tancredi, V., Biagini, G., D'Antuono, M., Louvel, J., Pumain, R. and Avoli, M. (2000) Spindle-like thalamocortical synchronization in a rat brain slice preparation. Journal of Neurophysiology 84: 1093–7.CrossRefGoogle Scholar
Tang, A. C., Bartels, A. M. and Sejnowski, T. J. (1997) Effects of cholinergic modulation on responses of neocortical neurons to fluctuating input. Cerebral Cortex 7: 502–9.CrossRefGoogle ScholarPubMed
Terreberry, R. R. and Neafsey, E. J. (1987) The rat medial frontal cortex projects directly to autonomic regions of the brainstem. Brain Research Bulletin 19: 639–49.CrossRefGoogle ScholarPubMed
Terzano, M. G., Parrino, L. and Spaggiari, M. C. (1988) The cyclic alternating pattern sequences in the dynamic organization of sleep. Electroencephalography and Clinical Neurophysiology 69: 437–47.CrossRefGoogle Scholar
Thierry, A. M., Gioanni, Y., Degenetais, E. and Glowinski, J. (2000) Hippocampo-prefrontal cortex pathway: anatomical and electrophysiological characteristics. Hippocampus 10: 411–19.3.0.CO;2-A>CrossRefGoogle ScholarPubMed
Thomson, A. M. (1988a) Inhibitory postsynaptic potentials evoked in thalamic neurons by stimulation of the reticularis nucleus evoke slow spikes in isolated rat brain slices. Neuroscience 25: 491–502.CrossRefGoogle Scholar
Thomson, A. M. (1988b) Biphasic responses of thalamic neurons to GABA in isolated rat brain slices. Neuroscience 25: 503–12.CrossRefGoogle Scholar
Thomson, A. M. (1997) Activity-dependent properties of synaptic transmission at two classes of connections made by rat neocortical pyramidal neurons in vitro. Journal of Physiology (London) 502: 131–47.CrossRefGoogle ScholarPubMed
Thomson, A. M. and Deuchars, J. (1997) Synaptic interactions in neocortical local circuits: dual intracellular recordings in vitro. Cerebral Cortex 7: 510–22.CrossRefGoogle ScholarPubMed
Thomson, A. M. and Morris, O. T. (2002) Selectivity in the inter-laminar connections made by neocortical neurones. Journal of Neurocytology 31: 239–46.CrossRefGoogle ScholarPubMed
Thomson, A. M., West, D. C., Hahn, J. and Deuchars, J. (1996) Single axon IPSPs elicited in pyramidal cells by three classes of interneurons in slices of rat neocortex. Journal of Physiology (London) 496: 81–102.CrossRefGoogle ScholarPubMed
Thomson, A. M., West, D. C., Wang, Y. and Bannister, A. P. (2002) Synaptic connections and small circuits involving excitatory and inhibitory neurons in layers 2–5 of adult rat and cat neocortex: triple intracellular recordings and biocytin labelling in vitro. Cerebral Cortex 12: 936–53.CrossRefGoogle ScholarPubMed
Tigges, J., Tigges, M., Cross, N. A.et al. (1982) Subcortical structures projecting to visual cortical areas in squirrel monkey. Journal of Comparative Neurology 209: 29–40.CrossRefGoogle ScholarPubMed
Tigges, J., Walker, L. C. and Tigges, M. (1983) Subcortical projections to the occipital lobe and parietal lobes of the chimpanzee brain. Journal of Comparative Neurology 220: 106–15.CrossRefGoogle ScholarPubMed
Timofeev, I. and Steriade, M. (1996) Low-frequency rhythms in the thalamus of intact-cortex and decorticated cats. Journal of Neurophysiology 76: 4152–68.CrossRefGoogle ScholarPubMed
Timofeev, I. and Steriade, M. (1997) Fast (mainly 30–100 Hz) oscillations in the cat cerebellothalamic pathway and their synchronization with cortical potentials. Journal of Physiology (London) 504: 153–68.CrossRefGoogle ScholarPubMed
Timofeev, I. and Steriade, M. (1998) Cellular mechanisms underlying intrathalamic augmenting responses of reticular and relay neurons. Journal of Neurophysiology 79: 2716–29.CrossRefGoogle ScholarPubMed
Timofeev, I. and Steriade, M. (2004) Neocortical seizures: initiation, development and cessation. Neuroscience 123: 299–336.CrossRefGoogle ScholarPubMed
Timofeev, I., Contreras, D. and Steriade, M. (1996) Synaptic responsiveness of cortical and thalamic neurons during various phases of slow oscillation in cat. Journal of Physiology (London) 494: 265–78.CrossRefGoogle ScholarPubMed
Timofeev, I., Grenier, F., Bazhenov, M., Sejnowski, T. J. and Steriade, M. (2000) Origin of slow oscillations in deafferented cortical slabs. Cerebral Cortex 10: 1185–99.CrossRefGoogle ScholarPubMed
Timofeev, I., Bazhenov, M., Sejnowski, T. J. and Steriade, M. (2001a) Contribution of intrinsic and synaptic factors in the desynchronization of thalamic oscillatory activity. Thalamus and Related Systems 1: 53–69.Google Scholar
Timofeev, I., Grenier, F. and Steriade, M. (2001b) Disfacilitation and active inhibition in the neocortex during the natural sleep-wake cycle: an intracellular study. Proceedings of the National Academy of Sciences USA 98: 1924–9.CrossRefGoogle Scholar
Timofeev, I., Bazhenov, M., Sejnowski, T. J. and Steriade, M. (2002a).Cortical IH takes part in the generation of paroxysmal activities. Proceedings of the National Academy of Sciences USA 99: 9533–37.CrossRefGoogle Scholar
Timofeev, I., Grenier, F., Bazhenov, M., Houweling, A., Sejnowski, T. J. and Steriade, M. (2002b).Short- and medium-term plasticity associated with augmenting responses in cortical slabs and spindles in intact cortex of cats in vivo. Journal of Physiology (London) 542: 583–98.CrossRefGoogle Scholar
Tomberg, C. (1999) Cognitive N140 electrogenesis and concomitant 40 Hz synchronization in mid-dorsal prefrontal cortex (area 46) identified in non-averaged human brain potentials. Neuroscience Letters 266: 141–4.CrossRefGoogle Scholar
Tomberg, C. and Desmedt, J. E. (1999) The challenge of non-invasive cognitive physiology of the human brain: how to negotiate the irrelevant background noise without spoiling the recorded data through electronic averaging. Philosophical Transactions of the Royal Society of LondonB354: 1295–305.CrossRefGoogle ScholarPubMed
Tömböl, T. and Szafranska-Kosmal, A. (1972) A Golgi study of the amygdaloid complex in the cat. Acta Neurobiologiae Experimentalis 32: 835–48.Google ScholarPubMed
Tononi, G. and Cirelli, C. (2001) Some considerations on sleep and neural plasticity. Archives Italiennes de Biologie 139: 221–41.Google ScholarPubMed
Tononi, G. and Cirelli, C. (2003) Sleep and synaptic homeostasis. Brain Research Bulletin 62: 143–50.CrossRefGoogle ScholarPubMed
Toth, T. L. and Crunelli, V. (1992) Computer simulation of the pacemaker oscillations of thalamocortical cells. NeuroReport 3: 65–8.CrossRefGoogle ScholarPubMed
Toth, T. I., Hughes, S. W. and Crunelli, V. (1998) Analysis and biophysical interpretation of bistable behaviour in thalamocortical neurones. Neuroscience 87: 519–23.CrossRefGoogle Scholar
Traub, R. D., Miles, R. and Wong, R. K. S. (1989) Model of the origin of rhythmic population oscillations in the hippocampal slice. Science 243: 1319–25.CrossRefGoogle ScholarPubMed
Traub, R. D., Buhl, E. H., Glovell, T. and Whittington, M. A. (2003) Fast rhythmic bursting can be induced by in layer 2/3 cortical neurons by enhancing persistent Na+ conductance and blocking BK channels. Journal of Neurophysiology 89: 909–21.CrossRefGoogle ScholarPubMed
Traub, R. D., Contreras, D., Cunningham, M. O.et al. (2005) Single-column thalamocortical network model exhibiting gamma oscillations, sleep spindles, and epileptogenic bursts. Journal of Neurophysiology 93: 2194–232.CrossRefGoogle ScholarPubMed
Trulson, M. E. and Jacobs, B. L. (1979) Raphe unit activity in freely moving cats: correlation with level of behavioral arousal. Brain Research 163: 135–50.CrossRefGoogle ScholarPubMed
Tsatsanis, K. D., Rourke, B. P., Klin, A.et al. (2003) Reduced thalamic volume in high-functioning individuals with autism. Biological Psychiatry 53: 121–9.CrossRefGoogle ScholarPubMed
Tseng, K. Y., Kasanetz, F., Kargieman, L., Riquelne, L. A. and Murer, M. G. (2001) Cortical slow oscillatory activity is reflected in the membrane potential and spike trains of striatal neurons in rats with chronic nigrostriatal lesions. Journal of Neuroscience 21: 6430–9.CrossRefGoogle ScholarPubMed
Tsvetkov, E., Carlezon, W. Jr., Benes, F. M., Kandel, E. R. and Bolshakov, V. Y. (2002) Fear conditioning occludes LTP-induced presynaptic enhancement of synaptic transmission in the cortical pathway to the lateral amygdala. Neuron 34: 289–300.CrossRefGoogle ScholarPubMed
Tsvetkov, E., Shin, R. M. and Bolshakov, V. Y. (2004) Glutamate uptake determines pathway specificity of long-term potentiation in the neural circuitry of fear conditioning. Neuron 41: 139–51.CrossRefGoogle ScholarPubMed
Turner, B. H. and Herkenham, M. (1991) Thalamoamygdaloid projections in the rat: a test of the amygdala's role in sensory processing. Journal of Comparative Neurology 313: 295–325.CrossRefGoogle ScholarPubMed
Turner, B. H. and Zimmer, J. (1984) The architecture and some of the interconnections of the rat's amygdala and lateral periallocortex. Journal of Comparative Neurology 227: 540–57.CrossRefGoogle ScholarPubMed
Uchida, S., Maloney, T., March, J. D., Azari, R. and Feinberg, I. (1991) Sigma (12–15 Hz) and delta (0.3–3.0) Hz EEG oscillate reciprocally within NREM sleep. Brain Research Bulletin 27: 93–6.CrossRefGoogle Scholar
Ulrich, D. and Huguenard, J. R. (1996) GAB AB-receptor-mediated responses in GABAergic projection neurones of rat nucleus reticularis thalami in vitro. Journal of Physiology (London) 493: 845–54.CrossRefGoogle Scholar
Uhlrich, D. J., Manning, K. A. and Pienkowski, T. P. (1993) The histaminergic innervation of the lateral geniculate complex in the cat. Visual Neuroscience 10: 225–35.CrossRefGoogle ScholarPubMed
Uhlrich, D. J., Manning, K. A. and Xue, J. T. (2002) Effects of activation of the histaminergic tuberomammillary nucleus on visual responses of neurons in the dorsal lateral geniculate nucleus. Journal of Neuroscience 22: 1098–107.CrossRefGoogle ScholarPubMed
Umbriaco, D., Watkins, K. C., Descarries, L., Cozzari, C. and Hartman, B. K. (1994) Ultrastructural and morphometric features of the acetylcholine innervation in adult rat parietal cortex. An electron microscopic study in serial sections. Journal of Comparative Neurology 348: 351–73.CrossRefGoogle ScholarPubMed
Urbain, N., Gervasoni, D., Soulière, F.et al. (2000) Unrelated course of subthalamic nucleus and globus pallidus neuronal activities across vigilance states in the rat. European Journal of Neuroscience 12: 3361–74.CrossRefGoogle ScholarPubMed
Urbain, N., Rentéro, N., Gervasoni, D., Renaud, B. and Chouvet, G. (2002) The switch of subthalamic neurons from an irregular to a bursting pattern does not solely depend on their GABAergic inputs in the anesthetic-free rat. Journal of Neuroscience 22: 8665–75.CrossRefGoogle Scholar
Ursin, H. and Kaada, B. R. (1960) Functional localization within the amygdaloid complex in the cat. Electroencephalography and Clinical Neurophysiology 12: 1–20.CrossRefGoogle ScholarPubMed
Uva, L., Gruschke, S., Biella, G., Curtis, M. and Witter, M. P. (2004) Cytoarchitectonic characterization of the parahippocampal region in guinea pigs. Journal of Comparative Neurology 474: 289–303.CrossRefGoogle Scholar
Brederode, J. and Spain, W. (1995) Differences in inhibitory synaptic input between layer II–III and layer V neurons of the cat neocortex. Journal of Neurophysiology 74: 1149–66.CrossRefGoogle Scholar
Hoesen, G. W. and Pandya, D. N. (1975) Some connections of the entorhinal (area 28) and perirhinal (area 35) cortices of the rhesus monkey. I. Temporal lobe afferents. Brain Research 95: 1–24.CrossRefGoogle ScholarPubMed
Vanni-Mercier, G. and Debilly, G. (1998) A key role for the caudoventral pontine tegmentum in the simultaneous generation of eye saccades in bursts and associated ponto-geniculo-occipital waves during paradoxical sleep in the cat. Neuroscience 86: 571–85.CrossRefGoogle ScholarPubMed
Veening, J. G., Swanson, L. W. and Sawchenko, P. E. (1984) The organization of projections from the central nucleus of the amygdala to brainstem sites involved in central autonomic regulation: a combined retrograde transport-immunohistochemical study. Brain Research 303: 337–57.CrossRefGoogle ScholarPubMed
Velayos, J. L., Jimenez-Castellanos, J. Jr. and Reinoso-Suárez, F. (1989) Topographical organization of the projections from the reticular thalamic nucleus to the intralaminar and medial thalamic nuclei in the cat. Journal of Comparative Neurology 279: 457–69.CrossRefGoogle ScholarPubMed
Venance, L., Rozov, A., Blatow, M., Burnashev, N., Feldmeyer, D. and Monyer, H. (2000) Connexin expression in electrically coupled postnatal rat brain neurons. Proceedings of the National Academy of Sciences USA 97: 10260–5.CrossRefGoogle ScholarPubMed
Vertes, R. P. (1991) A PHA-L analysis of ascending projections of the dorsal raphe nucleus in the rat. Journal of Comparative Neurology 313: 643–68.CrossRefGoogle ScholarPubMed
Vertes, R. P. (2004) Differential projections of the infralimbic and prelimbic cortex in the rat. Synapse 51: 32–58.CrossRefGoogle ScholarPubMed
Vertes, R. P., Fortin, W. J. and Crane, A. M. (1999) Projections of the median raphe nucleus in the rat. Journal of Comparative Neurology 407: 555–82.3.0.CO;2-E>CrossRefGoogle ScholarPubMed
Villablanca, J. (1965) The electrocorticogram in the chronic cerveau isolé cat. Electroencephalography and Clinical Neurophysiology 19: 576–86.CrossRefGoogle ScholarPubMed
Villablanca, J. (1974) Role of the thalamus in sleep control: sleep-wakefulness studies of chronic cats without the thalamus: the ‘athalamic cat’. In Basic Sleep Mechanisms, ed. Petre-Quadens, O. and Schlag, J., pp. 51–81. New York: Academic Press.Google Scholar
Vincent, S. R. and Reiner, P. B. (1987) The immunohistochemical localization of choline acetyltransferase in the cat brain. Brain Research Bulletin 18: 371–415.CrossRefGoogle ScholarPubMed
Vincent, S. R., Satoh, K., Armstrong, D. M. and Fibiger, H. C. (1983) Substance P in the ascending cholinergic reticular system. Nature 306: 688–91.CrossRefGoogle ScholarPubMed
Vincent, S. R., Satoh, K., Armstrong, D. M.et al. (1986) Neuropeptides and NADPH-diaphorase activity in the ascending cholinergic reticular system of the rat. Neuroscience 17: 167–82.CrossRefGoogle ScholarPubMed
der Malsburg, C. (1995) Binding in models of perception and brain function. Current Opinions in Neurobiology 5: 520–6.CrossRefGoogle Scholar
der Malsburg, C. (1999) The what and why of binding: the modeler's perspective. Neuron 24: 95–104.CrossRefGoogle Scholar
Krosigk, M., Bal, T. and McCormick, D. A. (1993) Cellular mechanisms of a synchronized oscillation in the thalamus. Science 261: 361–4.CrossRefGoogle Scholar
Wainer, B. H. and Mesulam, M.-M. (1990) Ascending cholinergic pathways in the rat brain. In Brain Cholinergic Systems, ed. Steriade, M. and Biesold, D., pp. 65–119. Oxford: Oxford University Press.Google Scholar
Wainer, B. H., Bolam, J. P., Freund, T. F.et al. (1984) Cholinergic synapses in the rat brain: a correlated light and electron microscopic immunohistochemical study employing a monoclonal antibody against choline acetyltransferase. Brain Research 308: 69–76.CrossRefGoogle ScholarPubMed
Walker, D. L. and Davis, M. (2000) Involvement of NMDA receptors within the amygdala in short- versus long-term memory for fear conditioning as assessed with fear-potentiated startle. Behavioral Neuroscience 114: 1019–33.CrossRefGoogle ScholarPubMed
Walker, D. L. and Davis, M. (2002) The role of amygdala glutamate receptors in fear learning, fear-potentiated startle, and extinction. Pharmacology Biochemistry and Behavior 71: 379–92.CrossRefGoogle ScholarPubMed
Walker, D. L., Ressler, K. J., Lu, K. T. and Davis, M. (2002) Facilitation of conditioned fear extinction by systemic administration or intra-amygdala infusions of D-Cycloserine as assessed with fear-potentiated startle in rats. Journal of Neuroscience 22: 2343–51.CrossRefGoogle ScholarPubMed
Wallace, D. M., Magnuson, D. J. and Gray, T. S. (1989) The amygdalo-brainstem pathway: selective innervation of dopaminergic, noradrenergic and adrenergic cells in the rat. Neuroscience Letters 97: 252–8.CrossRefGoogle ScholarPubMed
Wallace, D. M., Magnuson, D. J. and Gray, T. S. (1992) Organization of amygdaloid projections to brainstem dopaminergic, noradrenergic, and adrenergic cell groups in the rat. Brain Research Bulletin 28: 447–54.CrossRefGoogle ScholarPubMed
Wan, H., Warburton, E. C., Zhu, X. O.et al. (2004) Benzodiazepine impairment of perirhinal cortical plasticity and recognition memory. European Journal of Neuroscience 20: 2214–24.CrossRefGoogle ScholarPubMed
Wang, X. J. (1999) Fast burst firing and short-term synaptic plasticity: a model of neocortical chattering neurons. Neuroscience 89: 347–62.CrossRefGoogle ScholarPubMed
Wang, X. J. and Rinzel, J. (1993) Spindle rhythmicity in the reticularis thalami nucleus: synchronization among mutually inhibitory neurons. Neuroscience 53: 899–904.CrossRefGoogle ScholarPubMed
Wang, X. J., Tegnér, J., Constantinidis, C. and Goldman-Rakic, P. S. (2004) Division of labor among distinct subtypes of inhibitory neurons in a cortical microcircuit of working memory. Proceedings of the National Academy of Sciences USA 101: 1368–73.CrossRefGoogle Scholar
Wang, Z. and McCormick, D. A. (1993) Control of firing mode of corticotectal and corticopontine layer V burst-generating neurons by norepinephrine, acetylcholine and 1S, 3R-ACPD. Journal of Neuroscience 13: 2199–216.CrossRefGoogle Scholar
Warburton, E. C., Koder, T., Cho, K.et al. (2003) Cholinergic neurotransmission is essential for perirhinal cortical. Neuron 38: 987–96.CrossRefGoogle ScholarPubMed
Washburn, M. S. and Moises, H. C. (1992a) Electrophysiological and morphological properties of rat basolateral amygdaloid neurons in vitro. Journal of Neuroscience 12: 4066–79.CrossRefGoogle Scholar
Washburn, M. S. and Moises, H. C. (1992b) Inhibitory responses of rat basolateral amygdaloid neurons recorded in vitro. Neuroscience 50: 811–30.CrossRefGoogle Scholar
Washburn, M. S. and Moises, H. C. (1992c) Muscarinic responses of rat basolateral amygdaloid neurons recorded in vitro. Journal of Physiology (London) 449: 121–54.CrossRefGoogle Scholar
Watanabe, T., Taguchi, Y., Shiosaka, S.et al. (1984) Distribution of the histaminergic neuron system in the central nervous system of rats; a fluorescent immunohistochemical analysis with histidine decarboxylase as a marker. Brain Research 295: 13–25.CrossRefGoogle ScholarPubMed
Webster, H. H. and Jones, B. E. (1988) Neurotoxic lesions of the dorsolateral pontomesencephalic tegmentum cholinergic area in the cat. II. Effects upon sleep-waking states. Brain Research 458: 285–302.CrossRefGoogle ScholarPubMed
Weese, G. D., Phillips, J. M. and Brown, V. J. (1999) Attentional orienting is impaired by unilateral lesions of the thalamic reticular nucleus in the rat. Journal of Neuroscience 19: 10135–9.CrossRefGoogle ScholarPubMed
Weisskopf, M. G., Bauer, E. P. and LeDoux, J. E. (1999) L-type voltage-gated calcium channels mediate NMDA-independent associative long-term potentiation at thalamic input synapses to the amygdala. Journal of Neuroscience 19: 10512–19.CrossRefGoogle ScholarPubMed
Weliky, M. and Katz, L. C. (1999) Correlational structure of spontaneous neuronal activity in the developing lateral geniculate nucleus in vivo. Science 285: 599–604.CrossRefGoogle ScholarPubMed
Westgaard, F. H., Bonato, P. and Holte, K. A. (2000) Low-frequency oscillations (<0.3 Hz) in the electromyographic (EMG) activity of the human trapezius muscle during sleep. Journal of Neurophysiology 88: 1177–84.CrossRefGoogle Scholar
Weyand, T. G., Boudreaux, M. and Guido, W. (2001) Burst and tonic response modes in thalamic neurons during sleep and wakefulness. Journal of Neurophysiology 85: 1107–18.CrossRefGoogle ScholarPubMed
Whalen, P. J., Rauch, S. L., Etcoff, N. L.et al. (1998) Masked presentations of emotional facial expressions modulate amygdala activity without explicit knowledge. Journal of Neuroscience 18: 411–8.CrossRefGoogle ScholarPubMed
White, E. L. (1989) Cortical Circuits: Synaptic Organization of the Cerebral Cortex. Boston, MA: Birkhäuser.CrossRefGoogle Scholar
Wiedermann, U. A. and Lüthi, A. (2003) Timing of network synchronization by refractory mechanisms. Journal of Neurophysiology 90: 3902–11.CrossRefGoogle Scholar
Wilensky, A. E., Schafe, G. E. and LeDoux, J. E. (1999) Functional inactivation of the amygdala before but not after auditory fear conditioning prevents memory formation. Journal of Neuroscience 19: RC48.CrossRefGoogle Scholar
Wilensky, A. E., Schafe, G. E. and LeDoux, J. E. (2000) Functional inactivation of amygdala nuclei during acquisition of Pavlovian fear conditioning. Society for Neuroscience Abstracts 26: 465.Google Scholar
Wilensky, A. E., Schafe, G. E. and LeDoux, J. E. (2001) Does the central nucleus of the amygdala contribute to the consolidation of auditory fear conditioning. Society for Neuroscience Abstracts 27: 187.Google Scholar
Williams, J. A. and Reiner, P. B. (1993) Noradrenaline hyperpolarizes identified rat mesopontine cholinergic neurons in vitro. Journal of Neuroscience 13: 3878–83.CrossRefGoogle ScholarPubMed
Williams, J. A., Comisarow, J., Day, J., Fibiger, H. C. and Reiner, P. B. (1994) State-dependent release of acetylcholine in the rat thalamus measured by in vivo microdialysis. Journal of Neuroscience 14: 5236–42.CrossRefGoogle ScholarPubMed
Williams, J. T., North, R. A., Shefner, S. A., Nishi, S. and Egan, T. M. (1984) Membrane properties of rat locus coeruleus neurones. Neuroscience 13: 137–56.CrossRefGoogle ScholarPubMed
Williams, S. M. and Goldman-Rakic, P. S. (1998) Widespread origin of the primate mesofrontal dopamine system. Cerebral Cortex 8: 321–45.CrossRefGoogle ScholarPubMed
Williams, S. R., Toth, T. I., Turner, J. P., Hughes, S. W. and Crunelli, V. (1997) The ‘window’ component of the low threshold Ca2+ current produces input signal amplification and bistability in cat and rat thalamocortical neurones. Journal of Physiology (London) 505: 689–705.CrossRefGoogle ScholarPubMed
Wilson, C. J. (1987) Morphology and synaptic connections of crossed corticostriatal neurons in the rat. Journal of Comparative Neurology 263: 567–80.CrossRefGoogle ScholarPubMed
Wilson, C. J. and Kawaguchi, Y. (1996) The origin of two-state spontaneous membrane potential fluctuations of neostriatal spiny neurons. Journal of Neuroscience 16: 2397–410.CrossRefGoogle Scholar
Wilson, F. A. and Rolls, E. T. (1990) Learning and memory is reflected in the response of reinforcement-related neurons in the primate basal forebrain. Journal of Neuroscience 10: 1254–67.CrossRefGoogle ScholarPubMed
Wilson, M. A. and McNaughton, B. L. (1994) Reactivation of hippocampal ensemble memories during sleep. Science 265: 676–9.CrossRefGoogle ScholarPubMed
Wilson, M. A., Mascagni, F. and McDonald, A. J. (2002) Sex differences in delta opioid receptor immunoreactivity in rat medial amygdala. Neuroscience Letters 328: 160–4.CrossRefGoogle ScholarPubMed
Witter, M. P. and Groenewegen, H. J. (1984) Laminar origin and septotemporal distribution of entorhinal and perirhinal projections to the hippocampus in the cat. Journal of Comparative Neurology 224: 371–85.CrossRefGoogle ScholarPubMed
Witter, M. P. and Groenewegen, H. J. (1986) Connections of the parahippocampal cortex in the cat. IV. Subcortical efferents. Journal of Comparative Neurology 251: 51–77.CrossRefGoogle Scholar
Witter, M. P., Room, P., Groenewegen, H. J. and Lohman, A. H. M. (1986) Connections of the parahippocampal cortex in the cat. V. Intrinsic connections: Comments on input/output connections with the hippocampus. Journal of Comparative Neurology 252: 78–94.CrossRefGoogle ScholarPubMed
Witter, M. P., Groenewegen, H. J., Lopes da Silva, F. H. and Lohman, A. H. M. (1989) Functional organization of the extrinsic and intrinsic circuitry of the parahippocampal region. Progress in Neurobiology 33: 161–253.CrossRefGoogle ScholarPubMed
Witter, M. P., Wouterlood, F. G., Naber, P. A. and Haeften, T. (2000) Anatomical organization of the parahippocampal-hippocampal network. Annals of the New York Academy of Sciences 911: 1–24.CrossRefGoogle ScholarPubMed
Woody, C. D., Gruen, E. and Wang, X. F. (2003) Electrical properties affecting discharges of units of the mid and posterolateral thalamus of conscious cats. Neuroscience 122: 531–9.CrossRefGoogle ScholarPubMed
Woolf, N. J., Eckenstein, F. and Butcher, L. L. (1984) Cholinergic systems in the rat brain: I. Projections to the limbic telencephalon. Brain Research Bulletin 13: 751–84.CrossRefGoogle ScholarPubMed
Woolf, N. J., Hernit, M. C. and Butcher, L. L. (1986) Cholinergic and non-cholinergic projections from the rat basal forebrain. Neuroscience Letters 66: 281–6.CrossRefGoogle ScholarPubMed
Wouterlood, F. G. (2002) Cell types, local connectivity, microcircuits, and distribution of markers. In The Parahippocampal Region, ed. Witter, M. P. and Wouterlood, F. G., pp. 61–84. Oxford: Oxford University Press.Google Scholar
Wouterlood, F., Mugnaini, E. and Nederlof, J. (1985) Projection of olfactory lobe efferents to layer I GABA-ergic neurons in the entorhinal area. Combination of anterograde degeneration and immunoelectron microscopy in rat. Brain Research 343: 283–96.CrossRefGoogle Scholar
Wouterlood, F. G., Härtig, W., Brückner, G. and Witter, M. P. (1995) Parvalbumin-immunoreactive neurons in the entorhinal cortex of the rat: Localization, morphology, connectivity and ultrastructure. Journal of Neurocytology 24: 135–53.CrossRefGoogle ScholarPubMed
Wouterlood, F. G., Van, D. J., Haeften, T. and Witter, M. P. (2000) Calretinin in the entorhinal cortex of the rat: distribution, morphology, ultrastructure of neurons, and colocalization with gamma-aminobutyric acid and parvalbumin. Journal of Comparative Neurology 425: 177–92.3.0.CO;2-G>CrossRefGoogle Scholar
Wu, J. Y., Guan, L. and Tsau, Y. (1999) Propagating activation during oscillations and evoked responses in neocortical slices. Journal of Neuroscience 19: 5005–15.CrossRefGoogle ScholarPubMed
Xiang, J. Z. and Brown, M. W. (1998) Differential neuronal encoding of novelty, familiarity and recency in regions of the anterior temporal lobe. Neuropharmacology 37: 657–76.CrossRefGoogle ScholarPubMed
Xiang, Z., Huguenard, J. R. and Prince, D. A. (1998) Cholinergic switching within neocortical inhibitory networks. Science 281: 985–8.CrossRefGoogle ScholarPubMed
Xu, L., Tripathy, A., Pasek, D. A. and Meissner, G. (1999) Ruthenium red modifies the cardiac and skeletal muscle Ca2+ release channels (ryanodine receptors) by multiple mechanisms. Journal of Biological Chemistry 274: 32680–91.CrossRefGoogle Scholar
Yajeya, J., Juan, A. D., Bajo, V. M.et al. (1999) Muscarinic activation of a non-selective cationic conductance in pyramidal neurons in rat basolateral amygdala. Neuroscience 88: 159–67.CrossRefGoogle ScholarPubMed
Yajeya, J., Fuente, A., Criado, J. M.et al. (2000) Muscarinic agonist carbachol depresses excitatory synaptic transmission in the rat basolateral amygdala in vitro. Synapse 38: 151–60.3.0.CO;2-K>CrossRefGoogle ScholarPubMed
Yamada, T., Kameyama, S., Fuchigami, Y.et al. (1988) Changes of short latency somatosensory evoked potentials in sleep. Electroencephalography and Clinical Neurophysiology 70: 126–36.CrossRefGoogle ScholarPubMed
Yang, C. R., Seamans, J. K. and Gorelova, N. (1996) Electrophysiological and morphological properties of layers V–VI principal pyramidal cells in rat prefrontal cortex in vitro. Journal of Neuroscience 16: 1904–21.CrossRefGoogle ScholarPubMed
Yen, C. T., Conley, M., Hendry, S. H. C. and Jones, E. G. (1985) The morphology of physiologically identified GABAergic neurons in the somatic sensory part of the thalamic reticular nucleus in the cat. Journal of Neuroscience 5: 2254–68.CrossRefGoogle ScholarPubMed
Yingling, C. D. and Skinner, J. E. (1977) Gating of thalamic input to cerebral cortex by nucleus reticularis thalami. In Attention, Voluntary Contraction and Event-Related Cerebral Potentials, ed. Desmedt, J., pp. 534–59. Basel: Karger.Google Scholar
Ylinen, A., Bragin, A., Nádasdy, Z.et al. (1995) Sharp wave-associated high-frequency oscillation (200 Hz) in the intact hippocampus: network and intracellular mechanisms. Journal of Neuroscience 15: 30–46.CrossRefGoogle Scholar
Young, M. P., Tanaka, K. and Yamane, S. (1992) On oscillating neuronal responses in the visual cortex of the monkey. Journal of Neurophysiology 67: 1464–74.CrossRefGoogle ScholarPubMed
Yu, Y. Q., Xiong, Y., Chan, Y. S. and He, J. (2004) Corticofugal gating of auditory information in the thalamus: an in vivo intracellular recording study. Journal of Neuroscience 24: 3060–9.CrossRefGoogle Scholar
Yuste, R. and Tank, D. W. (1996) Dendritic integration in mammalian neurons, a century after Cajal. Neuron 16: 701–16.CrossRefGoogle ScholarPubMed
Zhang, L. and Jones, E. G. (2004) Corticothalamic inhibition in the thalamic reticular nucleus. Journal of Neurophysiology 91: 759–66.CrossRefGoogle ScholarPubMed
Zhang, S. J., Huguenard, J. R. and Prince, D. A. (1997) GAB AA receptor-mediated Cl− currents in rat thalamic reticular and relay neurons. Journal of Neurophysiology 78: 2280–6.CrossRefGoogle Scholar
Zhu, J. J. and Heggelund, P. (2001) Muscarinic regulation of dendritic and axonal outputs of rat thalamic interneurons: a new cellular mechanism for uncoupling distal dendrites. Journal of Neuroscience 21: 1148–59.CrossRefGoogle ScholarPubMed
Zhu, J. J. and Lo, F. S. (1999) Three GABA receptor-mediated postsynaptic potentials in interneurons in the rat lateral geniculate nucleus. Journal of Neuroscience 19: 5721–30.CrossRefGoogle ScholarPubMed
Zhu, J. J. and Uhlrich, D. J. (1998) Cellular mechanisms underlying two muscarinic receptor-mediated depolarizing responses in relay cells of the rat lateral geniculate nucleus. Neuroscience 87: 767–81.CrossRefGoogle ScholarPubMed
Zhu, J. J., Lytton, W. W., Xue, J. T. and Uhlrich, D. J. (1999a) An intrinsic oscillation in interneurons of the rat lateral geniculate nucleus. Journal of Neurophysiology 81: 702–11.CrossRefGoogle Scholar
Zhu, J. J., Uhlrich, D. J. and Lytton, W. W. (1999b) Burst firing in identified rat geniculate interneurons. Neuroscience 91: 1445–60.CrossRefGoogle Scholar
Zhu, J. J., Uhlrich, D. J. and Lytton, W. W. (1999c) Properties of a hyperpolarization-activated cation current in interneurons in the rat lateral geniculate nucleus. Neuroscience 92: 445–57.CrossRefGoogle Scholar
Ziakopoulos, Z., Tillett, C. W., Brown, M. W. and Bashir, Z. I. (1999) Input- and layer-dependent synaptic plasticity in the rat perirhinal cortex in vitro. Neuroscience 92: 459–72.CrossRefGoogle ScholarPubMed
Zola-Morgan, S., Squire, L. R., Amaral, D. G. and Suzuki, W. A. (1989) Lesions of perirhinal and parahippocampal cortex that spare the amygdala and hippocampal formation produce severe memory impairment. Journal of Neuroscience 9: 4355–70.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Mircea Steriade, Université Laval, Québec, Denis Pare, Rutgers University, New Jersey
  • Book: Gating in Cerebral Networks
  • Online publication: 18 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511541735.011
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Mircea Steriade, Université Laval, Québec, Denis Pare, Rutgers University, New Jersey
  • Book: Gating in Cerebral Networks
  • Online publication: 18 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511541735.011
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Mircea Steriade, Université Laval, Québec, Denis Pare, Rutgers University, New Jersey
  • Book: Gating in Cerebral Networks
  • Online publication: 18 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511541735.011
Available formats
×