Comprehensive Physiology Wiley Online Library

Neuronal Basis of Behavioral State Control

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Behavioral States and Biological Rhythms
1.1 State Concept and Definition of Some Behavioral States
1.2 Neuronal Correlates of Behavioral States and Their Interpretation as Causes
1.3 Temporal Aspects of Behavioral States
1.4 Circadian Rhythms and Behavioral State Control
2 Anatomical Substrates of Regulatory Systems
2.1 Changing the Concept of Nonspecificity
2.2 Brain Stem Reticular Formation
2.3 Thalamoneocortical Systems Related to Ascending Reticular Influences
2.4 Monoaminergic Systems
3 Brain Stem Core Systems Regulating Forebrain Activation
3.1 Definition of Activation
3.2 Search for Critical Regions at Global Level: Lesion‐Stimulation Studies
3.3 Criteria of Cellular Evaluation
3.4 Search for Candidate Mechanisms at Single‐Cell Level
4 Thalamocortical Mechanisms Related to Activation and Deactivation Processes
4.1 Spontaneous Activity
4.2 Excitatory‐Inhibitory Response Sequence
4.3 Excitability Enhancement During Attentional Tasks
4.4 Neurons and Transmitters Responsible for Thalamocortical Activation Processes
5 Sleep Cycle
5.1 Definition and Phenomenology
5.2 Cellular Correlates of Cycle
5.3 Pontine Localization of Desynchronized‐Sleep Trigger and Clock
5.4 Aminergic Hypothesis
5.5 Reciprocal Interaction Model of Sleep‐Cycle Control
5.6 Neuropharmacology of Sleep Cycle
6 Physiology and Pathophysiology of Sleep in Humans
6.1 Pathophysiological Model of Human Disorders
6.2 Pathophysiology of Narcolepsy
6.3 Pathophysiology of Sleep Apnea Syndromes
6.4 Disturbances of Motor Activity in Sleep
6.5 Dreaming and Disturbances of Mental Activity in Sleep
7 Functional Significance of Sleep and Waking States
7.1 General Adaptational Advantages of Rhythmic State Alternations
7.2 Rest Theory of Sleep
7.3 Development Implications of REM Sleep as Internal Activation Process
7.4 States and Acquisition of Learned Aspects of Adaptive Behavior
7.5 States and Metabolic Mode of the Brain
8 Conclusions
Figure 1. Figure 1.

Behavioral states in humans. States of waking, NREM sleep, and REM sleep have behavioral, polygraphic, and psychological manifestations. In behavior channel, posture shifts (detectable by time‐lapse photography or video) can occur during waking and in concert with phase changes of sleep cycle. [Two different mechanisms account for sleep immobility: disfacilitation (during stages I‐IV of NREM sleep) and inhibition (during REM sleep). In dreams, we imagine that we move but we do not.] Sequence of these stages represented in polygraph channel. Sample tracings of 3 variables used to distinguish state are also shown: electromyogram (EMG), which is highest in waking, intermediate in NREM sleep, and lowest in REM sleep; and electroencephalogram (EEG) and electrooculogram (EOG), which are both activated in waking and REM sleep and inactivated in NREM sleep. Each sample record is ∼20 s. Three lower channels describe other subjective and objective state variables.

Figure 2. Figure 2.

Behavioral states in cat. Polygraph records show that distinctive features of awake, NREM sleep, and REM sleep states shown in human records of Fig. 1 are shared by head‐restrained cat. The EMG, cortical EEG (EEGCtx), and EOG undergo sequence changes: progressive attentuation of muscle tone; activation, deactivation, and reactivation of cortical tone; and presence, absence, and reemergence of eye movement. Intracerebral leads reveal other features not detectable from surface recordings. The EEG of lateral geniculate body (EEGLGB) shows distinctively clustered biphasic waves that are synchronous with eye‐movement clusters of REM sleep; these waves are of considerably smaller amplitude in waking. Extracellular microelectrode recording (Cell) shows single‐cell action‐potential profiles of 3 types. Type A, most common, consists of lower discharge rate in NREM sleep than in waking or REM sleep; many neurons in cerebral cortex and cerebellum are type A. In type B progressively higher rates across the 3 states are seen in small proportion of cells, often motoneurons, especially those of pontine brain stem. Type C shows progressive decreases of rate across the 3 states, often with total cessation of discharge in REM sleep; this type of pattern, which is least common of the 3, is seen only in aminergic nuclei of brain stem. Each record is ∼25 s. (W. Silva and J. A. Hobson, unpublished observations.

Figure 3. Figure 3.

Circadian rhythms in humans. A: when human subject was isolated in underground bunker, period length of daily activity rhythm changed from 24 h after time cues were removed (day 3). This “free‐running” quality is cardinal characteristic of circadian rhythms and strongly suggests an endogenous origin. When time cues or zeitgebers are restored (day 21), rhythm was resynchronized to 24‐h period. B: 2 circadian rhythms may become dissociated from one another when both are allowed to run free. Sleep‐wakefulness rhythm has longer circadian period than body temperature rhythm. Thus more than 1 circadian clock must exist and be synchronized with one another by zeitgebers.

Adapted from Aschoff 40,41
Figure 4. Figure 4.

Circadian rhythms in rats. Circadian activity rhythms are released when normal animals, entrained to 24‐h zeitgebers (A), are deprived of light cues by blinding (B). These early records have been explained by discovery of retinal input to suprachiasmatic nucleus of hypothalamus.

Adapted from Richter 633
Figure 5. Figure 5.

Ultradian sleep cycle of NREM and REM sleep shown in detailed sleep‐stage graphs of 3 human subjects (A) and REM sleep periodograms of 15 human subjects (B). In polysomnograms of A, note typical preponderance of deepest stages (III and IV) of NREM sleep in the first 2 or 3 cycles of night; REM sleep is correspondingly brief (subjects 1 and 2) or even aborted (subject 3). During the last 2 cycles of night, NREM sleep is restricted to lighter stage (II), and REM periods occupy proportionally more of the time with individual episodes often exceeding 60 min (all 3 subjects). Same tendency to increase REM sleep duration is seen in B. In these records, all of which begin at sleep onset, not clock time, note variable latency to onset of first (usually short) REM sleep epoch. Thereafter inter‐REM period length is relatively constant. For both A and B time is in hours.

F. Snyder and J. A. Hobson, unpublished observations
Figure 6. Figure 6.

Biological rhythms and brain stem clocks. Three rhythms interact to determine cyclic order of sleep and waking states. Circadian rhythms are endogensus fluctuations of many bodily functions, including rest and activity, with periods of ∼24 h. As seen in schematic sagittal brain sections, suprachiasmatic nucleus of hypothalamus is key part of this control system that serves to synchronize internal processes with external forces. Ultradian sleep cycle, with its 90‐ to 100‐min period of NREM and REM sleep, is one of the physiological functions whose expression is circadian. It is controlled by reciprocal interaction of cholinergic and aminergic pontine reticular neurons, which oscillate out of phase with one another. This clock determines behavioral state (wake, NREM sleep, and REM sleep) of the organism. Mechanism by which circadian clock sets threshold of sleep‐cycle clock is unknown. Many homeostatic regulatory functions, including respiration, are influenced by circadian rhythm and sleep‐waking cycle. Respiratory oscillator is similar in neuronal design to sleep‐cycle clock but has shorter period (3 s) determined by reciprocal inhibition of expiratory and inspiratory neurons in medulla.

Figure 7. Figure 7.

Afferent projections to midbrain reticular formation (MRF) of cat. A: retrogradely labeled neurons in thalamic, subthalamic, and hypothalamic structures (2) after horseradish peroxidase injection into MRF (1). SC, superior colliculus; CG, central gray; RN, red nucleus; SN, substantia nigra; PP, pes pedunculi; LGd and LGv, dorsal and ventral lateral geniculate; OT, optic tract; VB, ventrobasal complex; PUL, pulvinar; CM‐PF, centrum medianum‐parafascicularis complex; RFB, retroflex bundle; FF, forel field; ZI, zona incerta; MTB, mamillothalamic bundle. B: reciprocal connections between MRF (recording) and CM‐PF (stimulation). Antidromic field (f) responses followed by synaptically elicited unit (u) discharges. Arrowhead, stimulus. Antidromic field response could follow 3 shocks at 250/s; its graded character is revealed by progressively decreasing stimulation intensity; unit discharges no longer appeared at lower intensity. C: latency histograms of synaptically (Syn) evoked discharges in MRF neurons to stimulation of CM, ZI, and preoptic area (POA). Coded neurons in histograms antidromically identified to project toward indicated sites. CL, centralis lateralis n. D: convergent synaptic excitation in MRF cell from bulbar (B) reticular formation and POA. Slow‐speed dotgrams (bottom) show dissimilar periods of suppressed firing after initial excitation induced by B and POA. E: graphs depicting percentage of MRF cells with various degrees of synaptic convergence in 2 neuronal populations (which could not be or have been antidromically identified from structures outside the MRF); 0 indicates neurons that have not been synaptically excited, and 1–4 indicate number of stimulated sites that induced synaptic excitation; relative segregation between nonprojection and projection elements in terms of degree of synaptic inputs is highly significant.

A adapted from Parent and Steriade 570 and Steriade et al. 760; B adapted from Steriade et al. 761; C and E adapted from Ropert and Steriade 647; D adapted from Steriade 738
Figure 8. Figure 8.

Ascending projections of rostral reticular formation in cat. A: drawings of selected parasagittal autoradiograms showing site of injection in cuneiform nucleus (bottom) and labeled projections ascending on ipsilateral side. Approximate laterality of sections in mm from median plane, ac, Anterior commissure; Bac, bed n. of anterior commissure; Bst, bed n. of stria terminalis; Ca, caudate n.; CD, central dorsal n.; CL‐PC, centralis lateralis‐paracentralis complex; dha, dorsal hypothalamic area; En, entopeduncular n.; IC, internal capsule; LD, laterodorsal n.; LP, lateralis posterior n.; lpa, lateral preoptic area; MD, mediodorsal n.; NPC, n. of posterior commissure; R, reticularis thalami n.; SI, substantia innominata; VL‐VA, ventralis lateralis‐ventralis anterior complex; VM, ventralis medialis n. B: diagrammatic chartings comparing patterns of anterograde labeling in 4 closely spaced transverse sections through intralaminar complex of thalamus (1) and mediodorsal n. (2) in cases of injections of tritium‐labeled amino acids into paramedian pontine tegmentum (1) and raphe‐interpeduncular complex (2). Hbl, habenula; NCM, n. centralis medialis. C: diagram representing percentages of antidromically identified MRF neurons from total number of tested elements. Depicted stimulating electrodes inserted into CM‐PF, CL, ZI, POA, paramedian pontine (P), and bulbar (B) reticular formation. Left: examples of CM‐evoked and ZI‐evoked antidromic discharges. S, collision with spontaneously occurring discharge.

A adapted from Edwards and De Olmos 205; B adapted from Graybiel 259; C adapted from Ropert and Steriade 647 and Steriade et al. 759
Figure 9. Figure 9.

Neocortical projections of intralaminar thalamic neurons and their monosynaptic excitation from midbrain reticular core in cat. A: calvarium with last recording thalamic microelectrode (Th) and chronically implanted stimulating electrodes in pericruciate motor cortex (M), parietal association cortex (P), and MRF. EOG and EEG, silver balls for recording eye movements and EEG rhythms; H, electrodes for recording hippocampal rhythms. B: lesion of ipsilateral pontine tegmentum for chronic degeneration of ascending systems coursing through MRF. BC, brachium conjunctivum; BP, brachium pontis; CS, n. raphe centralis superior; IC, inferior colliculus; LC, locus coeruleus; PG, pontine gray; RPO, n. reticularis pontis oralis. C: array of stimulating electrodes into MRF; most lateral electrode track was found in an anterior section. D: location of precruciate cortical stimulating electrodes within deep layers of medial parts of areas 8 and 6. Black dots, whole territory of pericruciate and anterior suprasylvian gyri (various cytoarchitectonic areas are indicated), covered with stimulating electrodes by changing their position from one experiment to another. E: location of 28 CL‐PC neurons found between anterior planes 9 and 10 and studied statistically for spontaneous and evoked activities in waking‐sleep states. Anti and Syn, antidromic and synaptic responses; CeM, n. centralis medialis; Pc, n. paracentralis; Rh, n. rhomboidalis; VPM, n. ventralis posteromedialis. Arrowheads indicate CL. F: physiological identification of CL‐PC neurons. Two different cells, antidromically activated from internal capsule (IC), motor cortex (MC), or parietal cortex (PC), and synaptically driven from MRF. Arrowheads, stimulus artifacts. 2: Only first stimulus of MC 3‐shock train at 250/s is marked; arrow, fractionation of antidromically elicited discharge to last stimulus in MC train. Collision between cortically elicited antidromic spikes and MRF‐evoked synaptic discharge shown in right superimposition (1) and in 10‐sweep sequence (2).

Adapted from Steriade and Glenn 750 and Glenn and Steriade 252
Figure 10. Figure 10.

Retrograde labeling of intralaminar thalamic neurons after neocortical injections of horseradish peroxidase (HRP) in rat. A: distribution of labeled cells in both intralaminar CL‐PC n. and LP n. after injection in parietal cortex. B: distribution of labeled cells in both intralaminar CL n. and VL n. after injection in precentral agranular motor area. Thalamic nuclei: AD, anterodorsal; AM, anteromedial; AV, anteroventral; Ce, central medial; MV, medioventral; Pom, medial division of posterior complex; Pva, anterior paraventricular; Pvp, posterior paraventricular; Sm, submedial; VMb, basal part of ventromedial.

Adapted from Jones and Leavitt 363
Figure 11. Figure 11.

Neocortical layer I projection from nucleus VM in cat. A: drawings of anterogradely and retrogradely labeled structures after ipsilateral HRP injection into VM. Dorsal view of cortex. Stippled area, cortical extension of intense granular reaction in layer I. In sagittal section (at ∼2.3 mm from midline), HRP reaction product indicative of anterogradely transported HRP in layer I marked by heavy line; heavy dashed line, liminal grain density. Dots, presence of retrogradely labeled cells in layer VI; density of dots was drawn in close proportion to density of labeled neurons, but number of dots is drastically lower than number of neurons. Light dashed lines, border between gray and white matter. Location of cytoarchitectonic areas indicated by numbers. F, fornix; G. Splen, splenial gyrus; IC, inferior colliculus; OB, olfactory bulb; S. cru, cruciate sulcus; V3, third ventricle. B: retrogradely labeled neurons in layer VI of area 6 (∼2.3 mm from midline); cru, cruciate sulcus; pre and post, precruciate and postcruciate gyri. Neurons taken at greater magnification are from crown of precruciate gyrus. Calibrations in mm. C: 1 and 2 show pattern of VM‐ and VL‐evoked cortical response in medial part of area 6 to single‐ and 10‐Hz shocks recorded at surface and depth of 1.0 mm. Note recruiting (initially surface‐negative, depth‐positive) and augmenting (initially surface‐positive, depth‐negative) responses to stimulation of VM and VL, respectively. 3: Suppression of VM‐evoked surface‐negative wave during cortex superfusion with Mn2+ to reversibly block synaptic transmission. Control waves in deep layers unaffected (not depicted; see details and Fig. 9 in ref. 251).

Adapted from Glenn et al. 251
Figure 12. Figure 12.

Distribution of brain stem monoamine neurons and ascending serotonergic pathways. A: drawings of transverse hemisections through brain stem of cat to illustrate distribution of monoamine‐containing neurons. Open circle, serotonergic cell bodies; filled circles, catecholaminergic cell bodies. B: representation of major organizational features of ascending serotonergic systems of rat brain as revealed by light‐microscope radioautography after intraventricular administration of tritiated serotonin. BO, bulbus olfactorius; CP, cerebral peduncle; CS, n. centralis superior; CT, corticospinal tract; DR, dorsal raphe n.; F, columna fornicis; FR, fasciculus retroflexus; GP, globus pallidus; GPO, griseum pontis; HI, hippocampus; IC, inferior colliculus; IP, interpeduncular n.; L, n. linearis rostralis; LL, n. of lateral lemniscus; MFB, medial forebrain bundle; MH, n. medialis habenulae; ML, medial lemniscus; MLF, medial longitudinal fasciculus; nVII, n. of facial nerve; PBC, n. parabrachialis; PVS, periventricular system; PY, pyramidal tract; RB, restiform body; RMA, n. raphe magnus; RP, n. raphe pallidus; SC, n. subcoeruleus; SL, n. septi lateralis; SM (caudal), stria medullaris thalami; SM (rostral), n. septi medialis; SNc, substantia nigra, pars compacta; SNr, substantia nigra, pars reticulata; TTS, transtegmental system; VTA, ventral tegmental area.

A adapted from Parent 568; B adapted from Parent et al. 569
Figure 13. Figure 13.

Early and late electrographic signs induced by kainic acid (KA) injection into MRF of chronically implanted, nonanesthetized cat and histological aspect of kainic‐induced lesion. A: the 3 ink‐written traces depict (before and after kainic injection) EEG waves, ocular movements (EOG), and EMG of neck muscles. Note normal fluctuations of EEG desynchronization‐synchronization periods before injection (top traces) and continuous EEG desynchronization (associated with highly aroused behavior) after injection. Picture (5 h after kainic injection) began ∼30 s after onset of injection and lasted for 24 h. This period corresponds to early excitation induced by KA. B: 2–4 days after kainic injection, corresponding to period of neuronal destruction, long‐lasting (∼30 s) EEG desynchronization elicited by an arousing (auditory) stimulus in control period before injection (top traces) is replaced by phasic EEG desynchronization, despite similar peripheral signs elicited by arousing stimulus. C: histological aspect of lesion produced by unilateral (left) injection (2.5 μg) of KA in MRF. Frontal plane 3.2. Animal was chronically implanted on both sides with guide cannulas into superior colliculi (SC) (vertical arrow in 1) and sacrificed 10 days after kainic injection. Territory of total neuronal loss on left side delimitated by dots (1). Limits of lesioned area may roughly indicate limits of initially excited area. Details at higher magnifications of left depicted in 2. Oblique arrow, blood vessel (permits localization of areas photographed in 2). Arrowheads, normal ganglion‐type cells of mesencephalic trigeminal nucleus, surviving within completely depopulated region (for details on selective and absolute resistance of these ganglion‐type cells, see refs. 154,177). Lesion affected almost entire MRF, leaving intact only a limited area dorsolateral to RN. Lesion encroached on lateral part of CG and lateral part of deep layers of SC. With exception of ganglion‐type trigeminal cells, there was total loss of neurons within lesioned area. Sections stained with unsuppressed Nauta method showed integrity of axons in fields with complete neuronal depopulation. FTC, central tegmental field; III, oculomotor nucleus.

Adapted from Kitsikis and Steriade 396 and Steriade 867
Figure 14. Figure 14.

Firing rates and patterns of MRF neurons in cat. A: discharge rate as function of dorsoventral localization. 1: Location at frontal plane 2.5 of neurons whose firing rates were roughly estimated during recording and divided into 3 classes: silent, <3/s, and >3/s. AQ, aqueduct; 3, oculomotor n.; L, raphe linearis. 2: Dorsoventral position of 44 MRF neurons whose firing rates during quiet waking were measured in detail; Spearman rank correlation between high discharge rate and ventral location (ra = 0.42) significant at P < 0.005; arrow, limit between deep layers of SC and MRF. B: discharge patterns of rostrally projecting MRF cell, antidromically identified from zona incerta and synaptically driven from preoptic area and paramedian pons. Note sustained regular firing regardless of movements in waking (W+) or quiet wakefulness (W‐). ISIH, interspike interval histogram; N, number of intervals; X, mean interval; M, modal interval; C, variation coefficient; E, proportion of intervals in excess of depicted time range. Note symmetric shape and interval density near mode. Bottom: autocorrelations (ACFs) of same neuron during states of W‐, slow‐wave sleep (S), and desynchronized sleep (D). Note, during W‐, rhythmic firing with peaks at multiples of mode, and flat contours in S and D states.

Adapted from Steriade et al. 759
Figure 15. Figure 15.

Midbrain reticular formation neurons increase discharge rates in advance of behavioral and EEG signs of activated states in cat. A: electrographic criteria of transitional state (SW) from slow‐wave sleep (S) to wakefulness (W). Abrupt (1) and progressive transition, with an intermediate SW period (2). B: percent cumulative histograms (1‐s bins) of 2 MRF neurons (neuron 2 antidromically identified from CM‐PF). Abscissas, real time of recording; arrows and vertical lines, earliest signs of reduced amplitude and/or increased frequency of EEG waves (as in A, pt. 2, arrow indicates time 0 of SW period). Inflection points are seen to occur 10–22 s in advance of any change in EEG; overt signs of wakefulness (eye movements and increased muscular tone) appeared several seconds after arrows (as in A, pt. 2). C: increase in firing rate of MRF neurons of cat before end of S epochs developing into W. Left: percent cumulative histogram of neuron belonging to sample analyzed in graph depicted on right; arrows, first change in fully synchronized EEG waves. Right: 25 cells whose global mean rate in S was at least 4/s were analyzed during last minute of S in 42 epochs leading to SW or directly to W. Mann‐Whitney test was used to compare reference rate during first 30 s for all cells with their respective rates in the last 6 5‐s bins. Note significantly increased rates in the 3 5‐s bins before end of S (arrow) compared with discharge rate in the first 30 s.

Adapted from Steriade et al. 759,761
Figure 16. Figure 16.

Midbrain reticular formation neurons decrease discharge rates in advance of first electro‐graphic signs during transition from W to S. A: electrographic criteria of transitional state from W to S (WS). Graph shows median of firing rates in sample of 52 MRF cells during W, WS, and S; note that major and significant decrease in firing rate occurs from W to WS. B: neuron antidromically identified from CM‐PF. Top 2 traces, original spikes and EEG waves simultaneously displayed on oscilloscope. Note decreased firing rate, leading to neuronal silence, before first EEG spindle sequence (between arrows) during transition from W to S. Bottom traces, same activities during repeated EEG desynchronization‐synchronization transitions (polygraphic recordings). C: 12 transitions of unit firing with respect to start of EEG spindle (time 0). Arrow, level of discharge (median rate) during W. Asterisks, bins with significant (<0.05) decrease in firing rate compared with median rate during W. Significantly decreased firing rate occurred 1 s before spindle onset.

Adapted from Steriade 737,740 and Steriade et al. 759
Figure 17. Figure 17.

Cortical and thalamic neuronal activity during waking‐sleep states. A: unit discharges from cell in motor cortex of monkey during W and S. Surface EEG rhythms from motor and visual cortices shown below discharges. Note that cell remained active during S. B: patterns of firing of lateral geniculate cell in cat during W and S. Upper beam swept from below upward; sweeps triggered by spikes. Positive deflections upward for continuous beam, to left for swept beam. Different time calibrations for upper and lower beam. Note bursting discharges during S. C: discharge frequency in antidromically identified pyramidal tract (PT) neurons of monkey and in neurons that were not antidromically identified (non‐PT) during W, S, and D. Ordinates, spikes/s in PT and non‐PT cells. The PT and non‐PT cells are quite different with respect to both total amount of activity and changes in amount of activity as a function of sleep and waking.

A adapted from Jasper 341; B adapted from Hubel 324; C adapted from Evarts 209
Figure 18. Figure 18.

Discharge patterns of thalamocortical and corticothalamic neurons during waking‐sleep states in cat. A and B: intralaminar CL‐PC thalamic neurons antidromically identified from MC (neuron B was also synaptically activated from midbrain reticular core). Note high‐frequency spike clusters in S and their replacement by sustained discharge in both behavioral states of W and D. C: cortical cell recorded in area 5 and backfired from the CM thalamic n. (see inset at left with antidromic identification). Similarly sustained discharges in W and D; decreased firing rate in S. Bottom: polygraph traces depict unit spikes, focal EEG waves simultaneously recorded by microelectrode in area 5, EEG from depth of visual cortex, EMG of neck muscles, and EOG. Note tonically increased firing rate preceding onset of D (upward arrow) and during D, similar to discharge pattern in W (awakening marked by downward arrow). Two parts are separated by nondepicted period of 180 s. Tonic discharge of corticofugal neurons throughout D is very different from burst discharges of cortical interneurons related to REM epochs (see Fig. 20C).

A and B adapted from Glenn and Steriade 252; C adapted from Steriade 736
Figure 19. Figure 19.

Fast‐conducting and slow‐conducting PT neurons on arousal and subsequent steady waking in monkey and cat. A: histogram of antidromic response latencies of 67 precentral PT neurons after peduncular stimulation in monkey. Hatched columns, units in which spontaneous discharges stopped on arousal from synchronized sleep; white columns, neurons whose spontaneous firing increased on arousal. Inset: PT cells (a, b) that are represented in B. Arrowhead, stimulus; oblique arrow, lack of antidromic spike of cell b due to collision with spontaneous (S) discharge. B: 4 ink‐written traces depict a and b units, EEG waves, and eye movements. Note that fast‐conducting a neuron diminished firing rate on arousal but, afterward, during steady waking, increased its firing rate over value seen in synchronized sleep. Slow‐conducting b cell increased firing rate from beginning of arousal. C: intracellular recording of fast PT cell in midpontine pretrigeminal cat. Traces, from top to bottom: stimulation of MRF at 100/s, EEG from MC, hippocampal activity, zero potential level (broken line), cell activity (spike peaks off), spike rate/s, and cell activity at higher gain (spike peaks off). Bottom: antidromic responses to peduncular stimulation and spontaneous action potentials.

A and B adapted from Steriade et al. 747; C adapted from Inubushi et al. 329
Figure 20. Figure 20.

Discharge patterns of cortical interneurons during waking‐sleep states in cat. A: discharges during W, S, and D of putative short‐axon cell recorded from parietal association area 5, synaptically driven with high‐frequency spike barrage from LP thalamic n. (similar to interneuron depicted in C). B: similar type of cell in area 5. Polygraph traces, from top to bottom: unit discharges, EEG focal waves recorded by microelectrode, surface EEG waves, EMG, and ocular movements. Note diminution of spontaneous firing during arousal elicited by auditory stimulation (between arrows) and progressively increased occurrence of spike bursts with transition from W to S. C: another interneuron in area 5, driven by stimulation of LP thalamic n. with spike barrage at ∼500/s occurring in relation with onset of slow positive focal wave. Note stereotyped spontaneous spike bursts in all (W, S, and D) states. Note in polygraph traces (same activities as in B) that, during D (beginning at arrow), interneuronal discharge bursts are closely related to REM episodes and are dissimilar to sustained firing of corticofugal neurons during D, as shown in Fig. 18C.

Adapted from Steriade 736 and Steriade et al. 758
Figure 21. Figure 21.

Antidromic responsiveness of thalamocortical cells during waking‐sleep states in cat. A: inhibition during EEG spindles of antidromic responses in thalamocortical VL neurons. Two simultaneously recorded cells. Neuron b was superimposed on negative hump of neuron a; note that neuron b was antidromically activated in absence of discharge of neuron a (5). Figures on EEG record (1–6) indicate application of testing shock trains, corresponding to those in traces depicting evoked discharges. Abolition of antidromic responses occurs during EEG spindles (3 and 6). Evoked discharge of cell a disappeared 500 ms prior to EEG spindle (5). B, top: antidromic responses to area 5 stimulation in typical CL cell of this sample during W, S, and D. Arrowheads, stimuli. Bottom: median (arrowheads), mean (columns), and standard deviation for percentage of antidromic responsiveness (R) in sample of 19 CL‐PC intralaminar thalamic neurons backfired from pericruciate (areas 6 and 4) or anterior suprasylvian (area 5) gyri during W and S. Five of these neurons were also recorded during D. Asterisk, collision with spontaneous discharge.

A adapted from Steriade et al. 741; B adapted from Glenn and Steriade 252
Figure 22. Figure 22.

Antidromic responsiveness of corticofugal cells during waking‐sleep states in cat and monkey. A: facilitation of antidromic invasion of PT cell of cat during EEG desynchronization. Full responsiveness to a 4‐shock train during control period of spontaneous EEG desynchronization (1), depressed responsiveness during progressively developing EEG synchronization (2 and 3), and recovery of antidromic responsiveness during EEG desynchronization elicited by brief conditioning pulse train to MRF (4). B, top: corticothalamic cell in area 5 of cat, antidromically activated from n. CM. Antidromic responsiveness was investigated with a 4‐shock train. Arrowheads, stimuli. Bottom: percentage responsiveness (R) to 1st, 2nd, and 4th shock during W, S, and D. Below each state is mean rate of spontaneous firing (during another waking‐sleep cycle). C: pattern of antidromic invasion of pre‐central PT neurons during behavioral S (left) and arousal (right) in monkey. Dotted line and arrow indicate arousal. Fast‐conducting neuron has 0.5‐ms antidromic response latency to peduncular stimulation. 1 and 2, Responses to trains of 5 shocks (350/s) and 3 shocks (110/s), respectively, during 2 passages from sleep to arousal. Third trace: EEG desynchronization. Bottom trace: eye movements. Note spike fragmentation of first antidromic spike (arrowheads) during sleep and full recovery on arousal as well as diminution of spike fragmentation of successive responses.

A adapted from Steriade 735; B adapted from Steriade et al. 754; C adapted from Steriade et al. 747
Figure 23. Figure 23.

Primary synaptic excitation in thalamus and neocortex during waking‐sleep states in cat. Evoked potential studies (A‐C) and extracellular unit recordings (D‐F). Filled circles, testing shocks. A: simultaneous recording of field potentials evoked in lateral geniculate n. (LG) and at surface of visual cortex (VC) by optic tract stimulations (monopolar recordings). Response of LG consists of presynaptic [tract (t)] positive component and monosynaptically relayed (r) negative component. Different components of VC response numbered 1–5. Note, during 300/s stimulation of MRF, enhancement of monosynaptic (r) component in LG without alteration in presynaptic (t) deflection; note also increased amplitude of VC response. B: surface VC potentials evoked by stimulation of white matter beneath VC (1) or to deep layers in VC (2). Note that, with white matter stimulation, postsynaptic components of VC response are enhanced without changes in presynaptic component. C: simultaneous recording of field potentials evoked in the VL thalamic n. and MC by stimulation of cerebellothalamic pathway during W, S, and D. Note in S, compared with both W and D, obliteration of monosynaptic thalamic wave (r) and marked reduction of cortical response, without alteration of component t that reflects activity in afferent fibers to VL n. (in this case, bipolar recording in VL). D: MRF‐evoked synaptic discharges in intralaminar CL thalamic neuron antidromically identified from parietal association area 5. Note decreased latency, increased probability of discharges in early bins, and shorter duration of spike bursts during W compared with S. E and F: unit recordings in SI cortex (SC) and white matter stimulation in animals with complete lesion of VB thalamic complex. The MRF shock train preceded testing shocks by 3–5 ms (not depicted). Note MRF‐induced facilitation of evoked discharges. F: MRF potentiation is seen by decrease in latency of discharges to first stimulus and increased discharge probability to second stimulus.

A and B adapted from Steriade 733,734; C adapted from Steriade et al. 752; D adapted from Steriade and Glenn 750; E and F adapted from Steriade and Morin 756
Figure 24. Figure 24.

Effects of natural arousal and midbrain reticular stimulation on inhibitory processes in thalamus and neocortex of cat and monkey. A: periods of suppressed firing (a, b, and c) and postinhibitory rebound excitation of neuron recorded in rostrodorsal part of LP thalamic n. after single‐shock stimulation (arrowhead) of cortical area 5 in cat. Preceding shock train to MRF reduced duration of periods a and b and abolished silent period c. B: patterns of VL‐evoked events in precentral neuron of monkey during S and W. Note that first inhibitory period evoked by VL stimulus (dots) persists in W, but subsequent rhythmic inhibition‐rebound sequences seen during S are abolished in W on background of increased spontaneous discharge. C: method of testing recurrent inhibition acting on antidromic discharges elicited in cat PT neurons by pes peduncular (PP) stimulation. Midbrain lesion included medial lemniscus to avoid stimulation of afferent fibers. Conditioning volley (C) was delivered at 13 V [minimal voltage required to elicit inhibitory effects on testing (T) response induced by shock at 5 V, which is minimal voltage required to evoke 100% antidromic invasion]. At paired C‐T stimulation, complete inhibition of T response or spike fragmentation (arrow). Graph depicts much longer inhibition with 3 PP shocks than with single shock. With both conditioning procedures (1 PP and 3 PP), recovery was slower during S than during W. D: inhibition of synaptic discharges evoked by stimulation of posterior part of VL in precentral PT neuron of monkey. Left: field positive (inhibitory) wave evoked by first VL stimulus and facilitation (during W) of evoked discharges by second stimulus at 75‐ms interval toward end of inhibition. Right: percentage responsiveness of discharge evoked by first stimulus (time 0) and by second stimulus at 3 time intervals (15 ms, 27 ms, and 75 ms) during W and S.

A adapted from Steriade et al. 757; B adapted from Steriade et al. 748; C and D adapted from Steriade and Deschěnes 745
Figure 25. Figure 25.

Secondary excitation and incremental responses during waking‐sleep states in cat. A: focal slow waves recorded at depth of 0.5 mm in cortical area 5 after stimulation of rostrodorsal part of LP thalamic n. during W, S, and D (50 averaged sweeps). Note during S selective enhancement of second (b) depth‐negative component. B: responses of cortical SI neuron to VB thalamic stimulation during S and W. Note during W increased probability of VB‐evoked early discharge and suppression of late repetitive discharges. C: simultaneous recording of field potentials (1: 50 averaged sweeps) and unit discharges (2) at depth of 1 mm in suprasylvian area 5, evoked by 2 0.1‐s‐delayed stimuli to LP thalamic n.; stimuli indicated by dots in 1 and by vertical bars in 2. Secondary depth‐negative wave b associated with repetitive discharges is selectively enhanced at second stimulus. D: poststimulus histograms of augmenting responses in sample of 7 cortical SI neurons, elicited by white matter (WM) stimuli at frequency of 10/s in VB‐lesioned preparation. S, stimulus number; R, responsiveness (total number of discharge to 100 shocks). Note that augmentation elicited by S‐2 (the second 0.1‐sdelayed shock) consists of increased secondary excitation (10–15 ms, arrow) simultaneous to decreased probability of primary synaptic excitation. Effect of a conditioning stimulation of MRF consists of an increased probability of primary excitation and a change in pattern of augmenting response to second stimulus into a primary one. E: augmenting field potentials at depth of 0.7 mm in area 5 to 10/s stimulation of LP thalamic nucleus during W, S, and D. Arrowheads in A, B, C, and E indicate stimuli.

A and E adapted from Steriade 738; B adapted from Steriade 735; C adapted from Steriade 736; D adapted from Steriade and Morin 756
Figure 26. Figure 26.

Cortically evoked rhythmic hyperpolarization‐rebound sequences. Intracellular recording of VL thalamic relay cell, antidromically invaded from precruciate MC and monosynaptically driven from brachium conjunctivum (BC). Resting membrane potential: −55 mV. 1: Typical response sequence. 2 to 4: 40 Averaged sweeps. Note in 3 and 4, analyzed at same speed, powerful rhythmic activity within frequency of spindle waves (∼7.5/s) evoked by MC, contrasting with lack of such activity after BC stimulation. Arrowheads, MC stimulation; filled circle, BC stimulation.

Adapted from Steriade 740
Figure 27. Figure 27.

Effect of membrane potential on excitability of VL thalamic relay neuron in cat; intracellular recording. A and B: antidromic and orthodromic activation by MC and BC stimulation, respectively. C: direct stimulation of cell by current pulse. At resting potential (0 nA) current pulse was subthreshold for spike initiation. During injection of 1 nA continuous hyperpolarizing current (2), same pulse remained subthreshold but break was followed by slow decaying response on voltage trace. Finally, under 2 nA hyperpolarizing current (3), pulse triggered “slow spike” and burst of action potentials. D: postanodal exaltation phenomenon obtained by passing hyperpolarizing current pulses of increasing intensities at resting potential. Note stereotyped burst response in D, pt. 3 compared with C, pt. 3. E: traces, same phenomenon as in D but at higher gain and slower speed. Current intensities similar to those in D. E, pt. 2: slow spike can be seen in isolation. Upper calibration bars apply to A and B.

Adapted from Deschěnes et al. 182
Figure 28. Figure 28.

Effects of rostral reticular stimulation on intracellulary recorded inhibitory processes in thalamic neurons. A: excitatory postsynaptic potential (EPSP) and inhibitory PSP (IPSP) sequences in VL neuron of cat during 7/s midline thalamic stimulation (1) and marked attenuation of hyperpolarizing potentials and increase in cell discharges during simultaneous low‐frequency thalamic stimulation and high‐frequency brain stem reticular stimulation (2). Upper traces of 1 and 2, evoked responses at motor cortical surface. B: VL relay cell in cat. 1: MC stimulation (arrowheads) evokes antidromic discharge followed by rhythmic inhibitory‐rebound sequences. 2: Conditioning 320/s shock train to MRF leaves intact the cortically elicited early IPSP but abolishes late phase of hyperpolarization and following rebound‐inhibition sequence. 3: Effect of MRF stimulation alone. C: slow (∼0.1 Hz) thalamic rhythm of hyperpolarizing episodes and its suppression during MRF stimulation. 1: Ink‐written intracellular recordings of VL relay cell. Note oscillations within frequency range of spindles (∼7 Hz), consisting of phasic hyperpolarizations followed by rebounds, appearing during slow rhythm (−0.12 Hz) of hyperpolarizing episodes, each lasting 2.3 s. Postinhibitory rebounds consist of slow spike (see Fig. 27) superimposed by fast repetitive action potentials. 2: Suppression of slow rhythm of thalamic hyperpolarizations during MRF stimulation (brief shock train every second); note increased background firing that outlasted MRF stimulation period; first episode of hyperpolarization after MRF stimulation lasted only 1.3 s, compared with 2.3‐s duration of these episodes before MRF stimulation.

A adapted from Purpura et al. 612; B and C adapted from Steriade 740
Figure 29. Figure 29.

Activities of RE neurons during waking‐sleep states in cat. Two neurons recorded in rostral pole of nucleus reticularis thalami, monosynaptically driven from precruciate gyrus with high‐frequency spike barrages, as shown for neuron A in top poststimulus histogram. T, number of trials; X, mean latency (ms); M, latency mode (ms); C, coefficient of variation; R, responsiveness (total number of spikes to 100 stimuli during depicted time). Traces in A: unit activity, surface cortical EEG waves, focal EEG waves recorded from electrode pair that induced synaptic activation of the neuron, EMG, and time (1 s). Note spike bursts closely related to EEG spindles and tonically increased firing rate during arousal and sustained waking. B: peristimulus histograms depict activities evoked from precruciate cortex during S and W. Left histograms (2‐ms bins) show details of early excitation; right histograms (25‐ms bins) show late inhibition‐rebound phases. Note in left histograms that probability of early evoked discharges (first 2‐ms bin) doubles in W compared with S. In S evoked burst mostly extends within latency range of secondary excitation (∼15–40 ms). Also note (right histograms) 2 evoked rebound sequences in S, whereas such events are lacking in W (see similar phenomena in Figs. 24A and 28B for thalamic relay cells).

Adapted from Steriade 740
Figure 30. Figure 30.

Facilitation of monkey's parietal visual neurons by attentive fixation. A: comparison of responses of parietal light‐sensitive neurons to visual stimuli in no‐task and task modes. Absence of responses during no‐task mode compared with strong responses in task mode is obvious. Summing histograms: standard error of mean calculated for each bin of the histograms; value shown by dotted line. Corresponding bin pairs within the histograms tested for significant differences (t test); bin pairs marked with diamonds differed at 5% level of significance. Overall response in no‐task and task states compared in following way. Rate of impulse discharge in prestimulus period was subtracted from that in poststimulus period for each trial, and populations of remainders were tested for significant differences (P ≪K 0.05 required) and used to form facilitation ratio for each neuron. Ratios for neurons with significant differences plotted in B. Facilitation ratio is that between net increment in response evoked by light stimulus in state of interested fixation over that evoked by physically identical and retinotopically similar stimulus delivered in no‐task or intertrial states. Fifty‐one neurons showed ratios of ≥1.0, and of these, difference was significant at 5% level (t test) for 38; values plotted in histogram. Ratio was fractional for 4 neurons indicated at left: for them, response was significantly greater in no‐task state than during interested fixation.

Adapted from Mountcastle et al. 528
Figure 31. Figure 31.

Facilitatory action of acetylcholine (ACh) in cerebral cortex. A: intracellular record from neuron in motor area of cat shows delayed depolarizing effect and prolonged firing evoked by iontophoretic application of ACh (140 nA). B: brief and instantly reversible depolarization and strong firing of same neuron caused by short intracellular current injection (monitored on lower trace). C: same neuron depressed after treatment with dinitrophenol; no longer fired in response to application of ACh (as in A); however, during continued ACh application, identical intracellular current injection (cf. B) now induced particularly powerful and prolonged discharge. The ACh thus greatly facilitates and prolongs any depolarizing input received by same cell. D and E: magnitude and time course of changes in potential and resistance induced by ACh. Open circles, resting potential: note slow and prolonged depolarizing effect; open triangles, resting resistance: note marked increase in resistance synchronous with depolarization; closed symbols, corresponding data recorded during IPSPs: they show relatively little change, except some possible reduction of inhibitory effect. F: summary of probable mechanism of action of ACh. In contrast to situation at neuromuscular junction (and other sites of nicotinic cholinergic action) where depolarization is mediated by enhancement of Na+ permeability, muscarinic ACh action in cerebral cortex tends to reduce membrane K+ permeability, causing an increase in overall resistance and facilitating depolarizing effect of other depolarizing inputs, such as EPSPs, which probably act by increasing Na+ permeability.

Adapted from Krnjević et al. 409
Figure 32. Figure 32.

Eye movements (EM), EEG, systolic blood pressure (SBP), respiration (resp), pulse, and body movements (BM) in a 100‐min sample of uninterrupted sleep over successive minutes of typical sleep cycle. Entire interval from minute 242 to 273 is considered to be REM period, even though eye movements (heavy bars) are not continuous.

Adapted from Snyder et al. 722
Figure 33. Figure 33.

PGO waves and relation to REM sleep and eye‐movement direction. A: NREM‐REM transition showing PGO waves in LGB (types I and II). During transition periods from NREM to REM sleep, biphasic (PGO) waves in LGB first appear as large single events (type I waves). Waves become clustered and of diminished and decrementing amplitude (type II waves) as signs of REM sleep become more prominent: atonia (EMG), desynchronization of cortical EEG (Cx), hippocampal‐θ (HIP), and REMs (EOG). B: side‐to‐side alternation of primary waves. Once REM period is well established, PGO wave amplitude primacy alternates from one geniculate to the other according to lateral direction of eye movements. When there is rightward movement of the eyes (EOG‐R), corresponding PGO wave cluster is larger in right LGB (dots) than in left. Conversely when there is leftward movement of the eyes (EOG‐L) waves are larger in left LGB (dots).

Adapted from Nelson et al. 856
Figure 34. Figure 34.

Behavioral state and neuronal discharge rate. Mean rates of brain cell populations shown as function of behavioral state. A: D‐on cells. Most cells of the brain have higher rates in D and W than in S. Cbm, cerebellar Purkinje cells; VN, vestibular n; Thal, thalamic n.; PRF, pontine reticular formation; Ctx, cerebral cortex; Hypo, hypothalamus. B: D‐off cells. Minority of cells that decrease rate in D are all found in or near aminergic zones of pontine brain stem. Ret N, reticular n. subpopulation; RN, raphe n.; PN, peribrachial neurons.

Adapted from Steriade and Hobson 751
Figure 35. Figure 35.

Relation of pontine burst cell firing, PGO wave amplitude, and eye movement lateral direction. A: burst cell and waves. Recording of cell in peribrachial region of pons that fires 4 clusters of spikes, each of which precedes PGO waves in LGB by 10–20 ms. Geniculate ipsilateral to cell (LGB1) shows markedly greater wave amplitude than the contralateral geniculate (LGBc) in initial pair of series. Correlation between pontine burst cell activity and LGB wave amplitude is 0.99. B: efferent copy. Three‐way correlation between eye movement direction (EOG), geniculate wave amplitude (LGB), and pontine burst cell activity shown on drawings of ventral brain surface (above) and exemplified in oscilloscope tracings (below). When eye movement is toward same side as cell, wave amplitude is greater in LGB1 and vice versa.

Adapted from Nelson et al. 856
Figure 36. Figure 36.

Pontine localization of REM sleep generator: lesion evidence. A: prepontine transection of the brain (with forebrain ablation) is followed by alternation of state resembling waking and REM sleep. Note suppression of muscle tone (EMG) and clustered eye movements (EOG). The EEG of brain stem shows low‐amplitude PGO waves occurring with eye‐movement clusters. Data show that REM sleep clock and trigger neurons are retropontine. B: in isolated pons preparation, C1 spinal cord transection is added to prepontine brain stem transection. Atonia is eliminated but REMs and pontine PGO waves appear with 30‐min periodicity characteristic of REM sleep in cat. Data show that REM sleep clock and trigger must reside in lower brain stem.

Adapted from Matsuzaki 474
Figure 37. Figure 37.

Pontine mechanisms of REM sleep generation: D‐on cells and executive actions. A: identification. Reticular neurons projecting to spinal cord can be identified by antidromic activation from axons at lumbar spinal cord levels via chronically implanted stimulating electrodes while recording somatic action potentials in brain stem of head‐restrained but unanesthetized cats with moveable microelectrode. At paramedian pontine site shown in 1, cell was recorded that showed a fixed‐latency response (2), fast following (3), and collision (4) with orthodromic spikes. Cell was typical in that it showed lowest firing rate in waking (5), a slightly greater rate in S (6), and intense bursting prior to REM bursts of D (7). B: selectivity. Tendency to concentrate firing in D quantified as ratio of mean rate in that state to rate in W or S. As shown in 1‐min episodes of spontaneous firing for each of these cells in the 3 sites, selectivity was highest in pontine reticular formation (1), next highest in pontomesencephalic reticular formation (2), and lowest in such precerebellar nuclei as tegmental reticular nucleus (TRC) (3). C: phasic latency. Tendency to fire prior to eye movements quantified by plotting sequential and cumulative histograms as shown and averaging across populations. Longest phase leads (up to 300 ms) and greatest increases (up to 100% of firing) were seen in pontine reticular formation. D: tonic latency. Pontine reticular neurons showed statistically significant increases in firing rate as early as 3.5 min prior to D epochs. These tonic rate changes were longer in gigantocellular tegmental field (FTG) than any other neuronal group in brain stem or forebrain. Note that rate increase occurs in episodic increments suggesting that phasic activation waves may spatially and temporally summate as recruitment spreads within the reticular pool. E: periodicity. Spontaneous discharge of single FTG neuron in 10 h of continuous recording. Discharge peaks occur at ∼30‐min intervals coincident with each REM period. F: proportion of firings by 4 single FTG neurons averaged in successive intervals of repeated sleep cycles. Each cycle was time normalized by establishing duration of D and dividing it into 5 equal parts; the 10 min before (D‐10) and after (D+10) each D period were also examined. For each cycle, percentage of total number of firings was determined for each of 25 bins.

Adapted from Wysinski et al. 833, Hobson et al. 308, Hobson et al. 309, Pivik et al. 595, and McCarley and Hobson 480
Figure 38. Figure 38.

Pontine mechanisms of REM sleep generation: D‐off cells and permissive action. In contrast to discharge pattern and activity profile on D‐on cells found in pontine reticular formation is behavior of neurons in aminergic nuclei such as LC cells. A: anatomical location. Histological reconstruction (drawing) and computer plot (inset) of microelectrode penetration and recording site, in n. LC (CAE), of D‐off cell. FTG, pontine reticular formation; 7G, genu, of facial nerve; SA, striae acousticae; VSL, lateral vestibular nucleus. B: spike trains. The LC cell recorded at site shown in A fired regularly in W, intermittently in S, and arrested discharge in D. Cell resumed firing, with no change in spike height, during subsequent cycle. C: dotgrams. One‐min periods of spontaneous activity in each of the 3 states shown by triggering Z‐axis of oscilloscope with action potentials shown in B. D: cell activity curves. As in D‐on cells, change in rate across states is progressive and continuous in D‐off cells but opposite in direction. Resumption of firing prior to end of D‐sleep epochs shown in insets of both graphs. SWS, slow‐wave sleep; PS, paradoxical sleep.

A, B, and C from J. A. Hobson, unpublished observations. D adapted from Aston‐Jones and Bloom 46
Figure 39. Figure 39.

A: reciprocal activity of LC and giant reticular neurons during sleep cycle. Reciprocal discharge by cells in CAE 559,560,561,562 and FTG 563,564,565,566,567. Outline tracing, sagittal section of cat pontine brain at 2.5 mm lateral to midline, showing path of an exploring microelectrode passing through cerebellum into dorsal brain stem. Inset, location of the 7 successive recording sites in penetration; circle diameter is 5 mm. Cumulative discharge histograms of cells recorded at 7 sites shown in inset during transition period beginning 2 min after D onset (vertical line) and ending 1 min thereafter. Each activity curve shows cumulative percentage of discharge for as much of the epoch as was free of arousal. Units recorded in LC and subcoeruleus decrease discharge rate (negatively inflected histograms), whereas those in FTG increase discharge rate (positively inflected histograms) with approach and onset of D. B: pooled geometric mean rate for population of 21 LC and 34 FTG neurons. Abscissa is W, S, and D. C: selectivity ratios of geometric mean rate for pooled population data shown in B. D: pooled cumulative histograms of discharge rates in subsets of the 2 neuronal groups during transition period from S to D. Ordinate, percentage of discharge; abscissa, time in minutes, where to is onset of D and 2 min before and 1 min after are shown. Overall mirroring of the 2 curves suggests that activity of the population may be modulated by shared, perhaps mutual, process.

Adapted from Hobson et al. 310
Figure 40. Figure 40.

Reciprocal interaction model of REM sleep generation. A: anatomical maps. Locus of D‐on cells (hatched area) in reticular formation nuclei and D‐off cells (dashed line) in pontine aminergic nuclei is shown on sagittal section of brain stem. Four sites with neurons relatively indifferent to state are shown on whole brain schematic: cerebral cortex (Ctx), cerebellar Purkinje cells (Cbm), other reticular nuclei (Ret), and PG. CAE, locus coeruleus; FTP, L, posterior and lateral tegmental fields; Pbl, parabrachial zone; RN, raphe n. B: selectivity gradient. The D/W selectivity ratios, computed and pooled as described in Fig. 39, are plotted for some of nuclei shown in A. Continuous gradient from most positively selective cells in FTG to most negatively selective cells in RN. Converse gradient would describe propensity to fire in waking under same recording conditions. C: physiological model. D‐on cells are assumed to be cholinergic/excitatory (E); D‐off cells are assumed to be aminergic/inhibitory (I). Each group is assumed to have both feedforward and feedback interconnections giving structural and synaptic model of reciprocal interaction. Model can be described mathematically and tested pharmacologically.

Figure 41. Figure 41.

Reciprocal interaction model: physiological and mathematical aspects. A: structural model of interaction between FTG and LC cell populations. Plus sign implies excitatory influences; minus sign implies inhibitory influences; a‐d correspond to constants associated with strength of connection and included in text equations. B: theoretical curve derived from model that best fits FTG unit in Fig. 37E. In this fit a = 0.3029, c = 0.1514, and the initial conditions (amplitude unsealed) were x(0) = 1 and y(0) = 4.5. C: histogram is average discharge level of another FTG unit over 12 sleep‐waking cycles, each normalized to constant duration. Cycle begins with end of D. Arrow, bin with most probable time of D onset. Solid curve, FTG fit; dotted curve, LC fit derived from model with values a 0.5490, c = 0.2745, x(0) = −1, and y(0) = 3.0 for the 2 populations. Dot, equilibrium value for the 2 populations. D: geometric mean values of discharge activity of 10 LC cells before (S), during (D), and after (W) a D episode. Each time epoch is equal to one‐quarter of D period. Note that discharge‐rate increase begins in last quarter of D, at same time that FTG activity is falling as predicted by model.

Figure 42. Figure 42.

Modulation of pontine neuronal excitability over sleep‐waking cycle. The finding that pontine reticular neurons fire in association with movement during waking in unrestrained cats is shown to be consequence of strong excitatory drive from other central motoneurons and afferent input from periphery. That different mechanisms could lead to same net output in W and D may be appreciated by comparing synaptic models (A) with activity pattern (B). In W powerful afferent excitation of FTG (++++) overcome tonic inhibitory restraint coming from aminergic neurons (—) and summates with recurrent collateral excitation (++) to produce clustered firing. In absence of afferent drive, cell is silent in W. In D less powerful afferent input (++) is capable of driving cells because level of inhibitory restraint has decreased (−). For same reason collateral excitation is more capable of sustaining firing within reticular pool. Result is repeated clustered firing that drives eye movements and muscle twitches. Because of tonic inhibition, motoneurons of large skeletal muscles do not respond; hence there is no major movement and no feedback from movement to central motor pattern generators in D. Because of phasic presynaptic inhibition of la cutaneous afferents there is also no change from peripheral sensory stimuli to drive system as is present in W. In slow‐wave sleep the system is in condition intermediate between the 2 extremes just discussed: disinhibition increases progressively and facilitation is markedly reduced so that cell firing is sparse.

Figure 43. Figure 43.

Pharmacology of REM sleep generation: comparison of results with predictions of reciprocal interaction model. REM sleep is favored (black postsynaptic neurons) by either enhancement of cholinergic transmission or by blockade of noradrenergic and serotonergic neurotransmission. Top: REM sleep increases are seen when neurotransmission is directly enhanced by agonists such as carbachol, which binds and stimulates acetylcholine receptors; REM sleep is also enhanced by eserine, which prevents enzymatic breakdown of endogenously released acetylcholine by cholinesterase. In contrast, REM sleep decreases are seen when muscarinic acetylcholine receptor is competitively occupied by atropine. Middle: REM sleep is decreased by any drug action that increases effective amount of this neurotransmitter. Monoamine oxidase (MAO) is inhibited (MAOI), leading to increase in NE and, via an inhibition of REM generators, a decrease in REM. In contrast, when β‐receptors are blocked by propranolol, postsynaptic REM generator cells are disinhibited and REM sleep is increased. Bottom: reuptake of aminergic neurotransmitter (e.g., by amitriptyline) may also lead to effective increase in serotonergic inhibition of REM sleep. On other hand, when reserpine depletes amount of serotonin (and NE) in vesicles, there is less inhibitory restraint and REM sleep increases.

Adapted from Vivaldi et al. 805
Figure 44. Figure 44.

Cholinergic REM sleep: enhancement effects of carbachol. A: predictions of reciprocal interaction model. Under physiological conditions, REM sleep occurs when D‐on cells are exposed to high levels of cholinergic excitation and low levels of aminergic inhibition. Carbachol increases cholinergic excitation of D‐on cells and thereby tips balance of system toward REM sleep generation. B: polygraph records. Qualitative identity, in all channels of physiological and pharmacological D. In both cases, EMG is flat, indicating atonia; cortical EEG (CX) is desynchronized, indicating activation; there are PGO waves in EEG of LGB; there is θ‐activity in hippocampal EEG (HIP); and REMs are indicated by deflections of EOG. C: individual time course data. Compared with control (above) note onset (<10 min) of prolonged (>60 min) D episodes seen after each of 3 carbachol cannula placements in pontine reticular formation. D: pooled time course data. Consistent enhancement of D seen across carbachol trials and subjects. At 1 h, peak value of 75% is 4–5 times greater than control and the effect lasts for 3 h.

Figure 45. Figure 45.

Cholinergic REM sleep enhancement: ionotophoretic effects of carbachol. A: chemical microstimulation. Microiontophoretic drug delivery system. Head restraint implant placed stereotaxically so that glass microiontophoretic pipette can be directed at pontine brain stem under chronic recording conditions. Electrographic D with carbachol (D‐carb) can be produced with currents as low as 300 nA for 10 min, and action potentials of single giant cells can be recorded through same pipette before and after delivery of drug. B: microiontophoretically induced D‐carb. Electrographic signs after delivery of carbachol by passing current through glass micropipette indistinguishable from those of physiological D and from D‐carb induced in cannula diffusion experiments. Injection was histologically localized to rostral pole of paramedian pontine reticular formation. C: iontophoretically induced PGO waves. At site different from that of B, note stereotyped groups of 4–8 waves that begin to occur, in absence of other D signs, at 33–52 min and are still present at 13 h, 42 min postinjection. Data suggest that it may be possible to chemically microdissect neuronal ensembles responsible for D‐state components. D: cross‐correlation of iontophoretically induced PGO waves and cell activity. Average PGO waveform and activity of single cell cross‐correlated in tape‐recorded segments of data shown in C. Note identity of PGO waveform (upper trace) to that seen in physiological D. Lower trace: histogram in which peak of PGO waves is taken as time zero and frequency of unit firing is counted for 500 ms before and after wave peak. Unit activity shows sharp increase in 40 ms prior to wave peak. This neuron histologically localized to PGO burst cell zone (see Fig. 35A).

Adapted from Vivaldi et al. 805
Figure 46. Figure 46.

Cholinergic REM sleep enhancement: intrabrain differentiation of enhancement and suppression stimulation sites. A: enhancement sites. Left: anterodorsal pontine reticular formation from which pure REM sleep is evoked. Middle: posteroventral pontine reticular formation from which REM sleep with stereotyped side‐to‐side alternation of eye movements (EMs) and PGO waves are evoked. Right: peribrachial pontine tegmentum from which state‐independent PGO waves are evoked. B: suppression sites. Left: midbrain reticular formation from which arousal with persistent motor circling is evoked. Middle: medullary reticular formation from which arousal with axial trunk rolling is evoked. Right: periabducens pontine tegmentum from which arousal with persistent oscillating eye movements (OEMs) is evoked. In each diagram parasagittal distance (in mm) is indicated. For sections at 1.2 mm: MLB, medial longitudinal bundle; TV, ventral tegmental n.; 6N, abducens nerve; TG, genu of facial nerve; 6, abducens nucleus; PH, praepositus hypoglossi. For section at 3.7: FTP, paralemniscal tegmental field; 5M, motor division of trigeminal nucleus; 7N, facial nerve; SO, superior olive; 7, facial nucleus; SA, striae acousticae; VL, lateral vestibular n.; S, solitary tract; VS, superior vestibular n.; EMG, nuchal EMG; EEG, cortical EEG; LGB, lateral geniculate body EEG; EOG, interorbital EOG. Each record ∼20 s.

H. Baghdoyan and J. A. Hobson, unpublished observations
Figure 47. Figure 47.

Cholinergic REM sleep enhancement: effects of bethanechol. A: molecular configuration. Top: acetylcholine, naturally occurring neurotransmitter that is rapidly degraded by cholinesterase. Middle: carbamylcholine (carbachol), long‐acting synthetic agent that resists enzymatic degradation. Carbachol is mixed nicotinic and muscarinic agonist. Bottom: β‐methyl carbamylcholine (bethanechol), long‐acting synthetic agent but pure muscarinic agonist. B: polygraphic records. EOG, flurries of REMs; EMG, complete suppression of nuchal muscle tone; EEG, desynchronization of cortical electrical activity; LGB, biphasic waves in LGB. Bethanechol‐induced state appears to be an intensified version of naturally occurring D state. C: time course data. Top: bethanechol trials and matched saline control recordings for 3 cats. At latencies of 10–30 min, D periods of up to 55 min ensue and are virtually continuous for 2 h. In all cases, latencies are shorter, durations longer, and intervals shorter than control values. Bottom: D‐beth values shown in top were pooled with 30‐min bins. Peak (85%) was reached at 1 h followed by progressive decline to trough (30%) at 3.5 h but with subsequent peaking. First time at which control levels matched experimental values was 5.5 h postinjection. D: anatomical differences indicate site specificity of enhancement and suppression of D by bethanechol. Enhancement of D was seen only at pontine sites where 4 of 7 trials produced at least a doubling of saline control values and only 1 produced suppression. In contrast, suppression was rule at medullary and midbrain sites where 6 of 8 trials produced at least a halving of control values and none produced enhancement. Four of the 6 medullary trials produced complete suppression of D.

M. Goldberg, E. Vivaldi, D. Riew, and J. A. Hobson, unpublished observations
Figure 48. Figure 48.

Antiaminergic REM‐sleep enhancement: effects of propranolol. A: predictions of reciprocal interaction model. Under physiological conditions, noradrenergic inhibition restrains cholinergic D generator. With β‐adrenergic blockade by microinjection or propranolol, D generator is disinhibited. B: polygraphic records show natural D onset and that seen after propranolol. There is no synchronized electroencephalographic activity in the cortex (CX) so that drug‐treated animal enters D directly from W. Channels as in Fig. 44. C: individual time course data. Enhancement of D by propranolol occurs by way of more episodes of normal length indicating that either threshold is not as effectively lowered as with carbachol and/or that serotonergic inhibition is capable of ending drug‐induced episodes. D: pooled time course data. Peak effect, at 1 h, is 2–3 times control. After return to base line at 3 h, there appears to be an undershoot or suppression that is not seen with cholinergic agonist enhancement.

Adapted from Vivaldi et al. 805
Figure 49. Figure 49.

Pathophysiology of narcolepsy. Narcoleptic patients have several abnormalities of sleep‐waking state control that appear to be result of changes in set point of REM state oscillator. During waking, REM sleep signs intrude as sleepiness, sleep attacks, and cataplexy. This increased propensity to REM is measured as decreases in multiple‐sleep latency test and may manifest itself in hypnogogic hallucinations or by sustained sleep‐onset REM period (normal subjects go quickly through stage I). On arousal from REM sleep, there may be corresponding persistence of mental phenomena of REM (hypnopompic hallucinations) and/or REM sleep motor inhibition (sleep paralysis). Reciprocal interaction model accounts for all these phenomena by hypothesizing decreased level of aminergic inhibition (light dashed line) and/or increased level of cholinergic excitation (light solid line) in pontine oscillator. During waking there is an increased propensity for REM generator to reach threshold. At sleep onset there is brief escape from aminergic restraint and sleep‐onset REM period is triggered. System then resets and cycles normally until end of REM when there is a lag in reinstitution of waking‐state conditions and REM phenomena again escape their normal temporal bounds. Clinical efficacy of aminergic agonists (e.g., amphetamine) or amine reuptake blockers (e.g., imipramine) may be due to their capacity to reset aminergic inhibition to normal level (heavy dashed line). By reciprocal interaction, this would also reset level of cholinergic generator (heavy solid line).

Figure 50. Figure 50.

Pathophysiology of sleep apnea syndrome. During waking, respiratory oscillator of medulla receives tonic drive from other neural structures and can respond to voluntary and metabolic signals to change breathing pattern. Ventilation is assured by active maintenance of airway pathway via tonus of oropharyngeal musculature. In NREM sleep, central drive on both respiratory oscillator and peripheral muscles declines due to disfacilitation. Respiratory rate and amplitude thus diminish and airway is subject to collapse. If obstruction occurs, forced expiratory effort may actually aggravate airway construction and prolonged apneas with marked hypoventilation and hypoxia may occur. During REM sleep, activation of pontine generator neurons produces tonic and phasic driving of respiratory oscillator, which may desynchronize leading to hyperpnea and/or to apnea. In addition, medullary oscillator becomes unresponsive to metabolic signals. In patients with tendency to airway collapse, these processes may multiply deleterious effects of ventilation.

Figure 51. Figure 51.

Motor activity in sleep: timing of posture shifts. A: posture shifts in video data. Movements are scored by making frame‐by‐frame comparisons of posture on replay of tape. Left, top and bottom: unambiguous posture shifts (>90% trunk rotation). Right, top and bottom: no such postural adjustments. Right, top: no movement is discernible; bottom: arm but no trunk movement. Drawings were made by tracing outlines of human subject in time‐lapse photographic study. B: individual records. Movement profiles characteristic of prompt (top) and delayed (bottom) sleep onset. Visual cross‐correlation between EEG data and posture shifts was established for each subject‐night. Movement ceases when NREM stages are progressive (top) but persists when waking and stage I alternate (bottom). Stereotyped coordination of posture shifts and sleep‐stage sequence followed sleep onset. Descending stages of NREM sleep, especially those reaching stages III and IV, were movement free. However, ascending NREM stages and REM (black areas) interruptions or terminations were accompanied by movements. C: pooled movement data. Immobility and sleep cycle phase. Top: average of 44 sleep cycles. Bottom: time of beginning (left curve) and ending (right curve) of 44 epochs of postural immobility related to average of all sleep cycles in which they occurred (top). Each curve is cumulative histogram of percentage of occurrences of immobility as function of percentage of cycle completed. Note steep and smoothly ascending curve of onsets indicating that immobility begins in association with early NREM sleep; curve of endings is by contrast inflected sharply at stage IV onset, which indicates that process controlling posture shifts is activated well before end of NREM sleep. D: neuronal activity curve. Actual activity of presumed REM generator neuron in cat brain stem (solid stepped curve) is compared with theoretical curves of on‐cell population (solid curve) and off‐cell population (dotted curve) as function of percent of cycle completed. Comparing D with C reveals inverse parallelism between timing curve of onsets in C of immobility and off‐cell trajectory in D; it is hypothesized that motor disfacilitation is occurring during first third of cycle. Later, curve of brain stem neuronal activation (D) directly parallels immobility offset in C. Both appear to be manifestations of phase shifts in motor systems such that immobility of REM sleep is produced by active peripheral inhibition in face of strong central facilitation.

Adapted from Aaronson et al. 1
Figure 52. Figure 52.

Mental activity in sleep: psychophysiology of dreaming. A: systems model. As a result of disinhibition caused by cessation of aminergic neuronal firing, brain stem reticular systems autoactivate. Their outputs have effects including depolarization of afferent terminals causing phasic presynaptic inhibition and blockade of external stimuli, especially during bursts of REM, and postsynaptic hyperpolarization causing tonic inhibition of motoneurons that effectively counteract concomitant motor commands so that somatic movement is blocked. Only oculomotor commands are read out as eye movements because motoneurons are not inhibited. Forebrain, activated by reticular formation and also aminergically disinhibited, receives efferent copy or corollary discharge information about somatic motor and oculomotor commands from which it may synthesize such internally generated perceptions as visual imagery and sensation of movement, both of which typify dream mentation. Forebrain may in turn generate its own motor commands, which help to perpetuate process via positive feedback to reticular formation. B: synaptic model. Some directly and indirectly disinhibited neuronal systems, together with their supposed contributions to REM sleep phenomena. At level of brain stem, 4 neuronal types are illustrated. MRF: midbrain reticular neurons projecting to thalamus that convey tonic and phasic activating signals rostrally; PGO: burst cells in peribrachial region that convey phasic activation and specific eye movement information to geniculate body and cortex (dashed line indicates uncertainty of direct projection); PRF: pontine reticular formation neurons that transmit phasic activation signals to oculomotor neurons (VI) and spinal cord, which generate eye movements, twitches of extremities, and presynaptic inhibition; BRF: bulbar reticular formation neurons that send tonic hyperpolarizing signals to motoneurons in spinal cord. As a consequence of these descending influences, sensory input and motor output are blocked at level of spinal cord. At level of forebrain, visual association and motor cortex neurons all receive time and phasic activation signals for nonspecific and specific thalamic relays.



Figure 1.

Behavioral states in humans. States of waking, NREM sleep, and REM sleep have behavioral, polygraphic, and psychological manifestations. In behavior channel, posture shifts (detectable by time‐lapse photography or video) can occur during waking and in concert with phase changes of sleep cycle. [Two different mechanisms account for sleep immobility: disfacilitation (during stages I‐IV of NREM sleep) and inhibition (during REM sleep). In dreams, we imagine that we move but we do not.] Sequence of these stages represented in polygraph channel. Sample tracings of 3 variables used to distinguish state are also shown: electromyogram (EMG), which is highest in waking, intermediate in NREM sleep, and lowest in REM sleep; and electroencephalogram (EEG) and electrooculogram (EOG), which are both activated in waking and REM sleep and inactivated in NREM sleep. Each sample record is ∼20 s. Three lower channels describe other subjective and objective state variables.



Figure 2.

Behavioral states in cat. Polygraph records show that distinctive features of awake, NREM sleep, and REM sleep states shown in human records of Fig. 1 are shared by head‐restrained cat. The EMG, cortical EEG (EEGCtx), and EOG undergo sequence changes: progressive attentuation of muscle tone; activation, deactivation, and reactivation of cortical tone; and presence, absence, and reemergence of eye movement. Intracerebral leads reveal other features not detectable from surface recordings. The EEG of lateral geniculate body (EEGLGB) shows distinctively clustered biphasic waves that are synchronous with eye‐movement clusters of REM sleep; these waves are of considerably smaller amplitude in waking. Extracellular microelectrode recording (Cell) shows single‐cell action‐potential profiles of 3 types. Type A, most common, consists of lower discharge rate in NREM sleep than in waking or REM sleep; many neurons in cerebral cortex and cerebellum are type A. In type B progressively higher rates across the 3 states are seen in small proportion of cells, often motoneurons, especially those of pontine brain stem. Type C shows progressive decreases of rate across the 3 states, often with total cessation of discharge in REM sleep; this type of pattern, which is least common of the 3, is seen only in aminergic nuclei of brain stem. Each record is ∼25 s. (W. Silva and J. A. Hobson, unpublished observations.



Figure 3.

Circadian rhythms in humans. A: when human subject was isolated in underground bunker, period length of daily activity rhythm changed from 24 h after time cues were removed (day 3). This “free‐running” quality is cardinal characteristic of circadian rhythms and strongly suggests an endogenous origin. When time cues or zeitgebers are restored (day 21), rhythm was resynchronized to 24‐h period. B: 2 circadian rhythms may become dissociated from one another when both are allowed to run free. Sleep‐wakefulness rhythm has longer circadian period than body temperature rhythm. Thus more than 1 circadian clock must exist and be synchronized with one another by zeitgebers.

Adapted from Aschoff 40,41


Figure 4.

Circadian rhythms in rats. Circadian activity rhythms are released when normal animals, entrained to 24‐h zeitgebers (A), are deprived of light cues by blinding (B). These early records have been explained by discovery of retinal input to suprachiasmatic nucleus of hypothalamus.

Adapted from Richter 633


Figure 5.

Ultradian sleep cycle of NREM and REM sleep shown in detailed sleep‐stage graphs of 3 human subjects (A) and REM sleep periodograms of 15 human subjects (B). In polysomnograms of A, note typical preponderance of deepest stages (III and IV) of NREM sleep in the first 2 or 3 cycles of night; REM sleep is correspondingly brief (subjects 1 and 2) or even aborted (subject 3). During the last 2 cycles of night, NREM sleep is restricted to lighter stage (II), and REM periods occupy proportionally more of the time with individual episodes often exceeding 60 min (all 3 subjects). Same tendency to increase REM sleep duration is seen in B. In these records, all of which begin at sleep onset, not clock time, note variable latency to onset of first (usually short) REM sleep epoch. Thereafter inter‐REM period length is relatively constant. For both A and B time is in hours.

F. Snyder and J. A. Hobson, unpublished observations


Figure 6.

Biological rhythms and brain stem clocks. Three rhythms interact to determine cyclic order of sleep and waking states. Circadian rhythms are endogensus fluctuations of many bodily functions, including rest and activity, with periods of ∼24 h. As seen in schematic sagittal brain sections, suprachiasmatic nucleus of hypothalamus is key part of this control system that serves to synchronize internal processes with external forces. Ultradian sleep cycle, with its 90‐ to 100‐min period of NREM and REM sleep, is one of the physiological functions whose expression is circadian. It is controlled by reciprocal interaction of cholinergic and aminergic pontine reticular neurons, which oscillate out of phase with one another. This clock determines behavioral state (wake, NREM sleep, and REM sleep) of the organism. Mechanism by which circadian clock sets threshold of sleep‐cycle clock is unknown. Many homeostatic regulatory functions, including respiration, are influenced by circadian rhythm and sleep‐waking cycle. Respiratory oscillator is similar in neuronal design to sleep‐cycle clock but has shorter period (3 s) determined by reciprocal inhibition of expiratory and inspiratory neurons in medulla.



Figure 7.

Afferent projections to midbrain reticular formation (MRF) of cat. A: retrogradely labeled neurons in thalamic, subthalamic, and hypothalamic structures (2) after horseradish peroxidase injection into MRF (1). SC, superior colliculus; CG, central gray; RN, red nucleus; SN, substantia nigra; PP, pes pedunculi; LGd and LGv, dorsal and ventral lateral geniculate; OT, optic tract; VB, ventrobasal complex; PUL, pulvinar; CM‐PF, centrum medianum‐parafascicularis complex; RFB, retroflex bundle; FF, forel field; ZI, zona incerta; MTB, mamillothalamic bundle. B: reciprocal connections between MRF (recording) and CM‐PF (stimulation). Antidromic field (f) responses followed by synaptically elicited unit (u) discharges. Arrowhead, stimulus. Antidromic field response could follow 3 shocks at 250/s; its graded character is revealed by progressively decreasing stimulation intensity; unit discharges no longer appeared at lower intensity. C: latency histograms of synaptically (Syn) evoked discharges in MRF neurons to stimulation of CM, ZI, and preoptic area (POA). Coded neurons in histograms antidromically identified to project toward indicated sites. CL, centralis lateralis n. D: convergent synaptic excitation in MRF cell from bulbar (B) reticular formation and POA. Slow‐speed dotgrams (bottom) show dissimilar periods of suppressed firing after initial excitation induced by B and POA. E: graphs depicting percentage of MRF cells with various degrees of synaptic convergence in 2 neuronal populations (which could not be or have been antidromically identified from structures outside the MRF); 0 indicates neurons that have not been synaptically excited, and 1–4 indicate number of stimulated sites that induced synaptic excitation; relative segregation between nonprojection and projection elements in terms of degree of synaptic inputs is highly significant.

A adapted from Parent and Steriade 570 and Steriade et al. 760; B adapted from Steriade et al. 761; C and E adapted from Ropert and Steriade 647; D adapted from Steriade 738


Figure 8.

Ascending projections of rostral reticular formation in cat. A: drawings of selected parasagittal autoradiograms showing site of injection in cuneiform nucleus (bottom) and labeled projections ascending on ipsilateral side. Approximate laterality of sections in mm from median plane, ac, Anterior commissure; Bac, bed n. of anterior commissure; Bst, bed n. of stria terminalis; Ca, caudate n.; CD, central dorsal n.; CL‐PC, centralis lateralis‐paracentralis complex; dha, dorsal hypothalamic area; En, entopeduncular n.; IC, internal capsule; LD, laterodorsal n.; LP, lateralis posterior n.; lpa, lateral preoptic area; MD, mediodorsal n.; NPC, n. of posterior commissure; R, reticularis thalami n.; SI, substantia innominata; VL‐VA, ventralis lateralis‐ventralis anterior complex; VM, ventralis medialis n. B: diagrammatic chartings comparing patterns of anterograde labeling in 4 closely spaced transverse sections through intralaminar complex of thalamus (1) and mediodorsal n. (2) in cases of injections of tritium‐labeled amino acids into paramedian pontine tegmentum (1) and raphe‐interpeduncular complex (2). Hbl, habenula; NCM, n. centralis medialis. C: diagram representing percentages of antidromically identified MRF neurons from total number of tested elements. Depicted stimulating electrodes inserted into CM‐PF, CL, ZI, POA, paramedian pontine (P), and bulbar (B) reticular formation. Left: examples of CM‐evoked and ZI‐evoked antidromic discharges. S, collision with spontaneously occurring discharge.

A adapted from Edwards and De Olmos 205; B adapted from Graybiel 259; C adapted from Ropert and Steriade 647 and Steriade et al. 759


Figure 9.

Neocortical projections of intralaminar thalamic neurons and their monosynaptic excitation from midbrain reticular core in cat. A: calvarium with last recording thalamic microelectrode (Th) and chronically implanted stimulating electrodes in pericruciate motor cortex (M), parietal association cortex (P), and MRF. EOG and EEG, silver balls for recording eye movements and EEG rhythms; H, electrodes for recording hippocampal rhythms. B: lesion of ipsilateral pontine tegmentum for chronic degeneration of ascending systems coursing through MRF. BC, brachium conjunctivum; BP, brachium pontis; CS, n. raphe centralis superior; IC, inferior colliculus; LC, locus coeruleus; PG, pontine gray; RPO, n. reticularis pontis oralis. C: array of stimulating electrodes into MRF; most lateral electrode track was found in an anterior section. D: location of precruciate cortical stimulating electrodes within deep layers of medial parts of areas 8 and 6. Black dots, whole territory of pericruciate and anterior suprasylvian gyri (various cytoarchitectonic areas are indicated), covered with stimulating electrodes by changing their position from one experiment to another. E: location of 28 CL‐PC neurons found between anterior planes 9 and 10 and studied statistically for spontaneous and evoked activities in waking‐sleep states. Anti and Syn, antidromic and synaptic responses; CeM, n. centralis medialis; Pc, n. paracentralis; Rh, n. rhomboidalis; VPM, n. ventralis posteromedialis. Arrowheads indicate CL. F: physiological identification of CL‐PC neurons. Two different cells, antidromically activated from internal capsule (IC), motor cortex (MC), or parietal cortex (PC), and synaptically driven from MRF. Arrowheads, stimulus artifacts. 2: Only first stimulus of MC 3‐shock train at 250/s is marked; arrow, fractionation of antidromically elicited discharge to last stimulus in MC train. Collision between cortically elicited antidromic spikes and MRF‐evoked synaptic discharge shown in right superimposition (1) and in 10‐sweep sequence (2).

Adapted from Steriade and Glenn 750 and Glenn and Steriade 252


Figure 10.

Retrograde labeling of intralaminar thalamic neurons after neocortical injections of horseradish peroxidase (HRP) in rat. A: distribution of labeled cells in both intralaminar CL‐PC n. and LP n. after injection in parietal cortex. B: distribution of labeled cells in both intralaminar CL n. and VL n. after injection in precentral agranular motor area. Thalamic nuclei: AD, anterodorsal; AM, anteromedial; AV, anteroventral; Ce, central medial; MV, medioventral; Pom, medial division of posterior complex; Pva, anterior paraventricular; Pvp, posterior paraventricular; Sm, submedial; VMb, basal part of ventromedial.

Adapted from Jones and Leavitt 363


Figure 11.

Neocortical layer I projection from nucleus VM in cat. A: drawings of anterogradely and retrogradely labeled structures after ipsilateral HRP injection into VM. Dorsal view of cortex. Stippled area, cortical extension of intense granular reaction in layer I. In sagittal section (at ∼2.3 mm from midline), HRP reaction product indicative of anterogradely transported HRP in layer I marked by heavy line; heavy dashed line, liminal grain density. Dots, presence of retrogradely labeled cells in layer VI; density of dots was drawn in close proportion to density of labeled neurons, but number of dots is drastically lower than number of neurons. Light dashed lines, border between gray and white matter. Location of cytoarchitectonic areas indicated by numbers. F, fornix; G. Splen, splenial gyrus; IC, inferior colliculus; OB, olfactory bulb; S. cru, cruciate sulcus; V3, third ventricle. B: retrogradely labeled neurons in layer VI of area 6 (∼2.3 mm from midline); cru, cruciate sulcus; pre and post, precruciate and postcruciate gyri. Neurons taken at greater magnification are from crown of precruciate gyrus. Calibrations in mm. C: 1 and 2 show pattern of VM‐ and VL‐evoked cortical response in medial part of area 6 to single‐ and 10‐Hz shocks recorded at surface and depth of 1.0 mm. Note recruiting (initially surface‐negative, depth‐positive) and augmenting (initially surface‐positive, depth‐negative) responses to stimulation of VM and VL, respectively. 3: Suppression of VM‐evoked surface‐negative wave during cortex superfusion with Mn2+ to reversibly block synaptic transmission. Control waves in deep layers unaffected (not depicted; see details and Fig. 9 in ref. 251).

Adapted from Glenn et al. 251


Figure 12.

Distribution of brain stem monoamine neurons and ascending serotonergic pathways. A: drawings of transverse hemisections through brain stem of cat to illustrate distribution of monoamine‐containing neurons. Open circle, serotonergic cell bodies; filled circles, catecholaminergic cell bodies. B: representation of major organizational features of ascending serotonergic systems of rat brain as revealed by light‐microscope radioautography after intraventricular administration of tritiated serotonin. BO, bulbus olfactorius; CP, cerebral peduncle; CS, n. centralis superior; CT, corticospinal tract; DR, dorsal raphe n.; F, columna fornicis; FR, fasciculus retroflexus; GP, globus pallidus; GPO, griseum pontis; HI, hippocampus; IC, inferior colliculus; IP, interpeduncular n.; L, n. linearis rostralis; LL, n. of lateral lemniscus; MFB, medial forebrain bundle; MH, n. medialis habenulae; ML, medial lemniscus; MLF, medial longitudinal fasciculus; nVII, n. of facial nerve; PBC, n. parabrachialis; PVS, periventricular system; PY, pyramidal tract; RB, restiform body; RMA, n. raphe magnus; RP, n. raphe pallidus; SC, n. subcoeruleus; SL, n. septi lateralis; SM (caudal), stria medullaris thalami; SM (rostral), n. septi medialis; SNc, substantia nigra, pars compacta; SNr, substantia nigra, pars reticulata; TTS, transtegmental system; VTA, ventral tegmental area.

A adapted from Parent 568; B adapted from Parent et al. 569


Figure 13.

Early and late electrographic signs induced by kainic acid (KA) injection into MRF of chronically implanted, nonanesthetized cat and histological aspect of kainic‐induced lesion. A: the 3 ink‐written traces depict (before and after kainic injection) EEG waves, ocular movements (EOG), and EMG of neck muscles. Note normal fluctuations of EEG desynchronization‐synchronization periods before injection (top traces) and continuous EEG desynchronization (associated with highly aroused behavior) after injection. Picture (5 h after kainic injection) began ∼30 s after onset of injection and lasted for 24 h. This period corresponds to early excitation induced by KA. B: 2–4 days after kainic injection, corresponding to period of neuronal destruction, long‐lasting (∼30 s) EEG desynchronization elicited by an arousing (auditory) stimulus in control period before injection (top traces) is replaced by phasic EEG desynchronization, despite similar peripheral signs elicited by arousing stimulus. C: histological aspect of lesion produced by unilateral (left) injection (2.5 μg) of KA in MRF. Frontal plane 3.2. Animal was chronically implanted on both sides with guide cannulas into superior colliculi (SC) (vertical arrow in 1) and sacrificed 10 days after kainic injection. Territory of total neuronal loss on left side delimitated by dots (1). Limits of lesioned area may roughly indicate limits of initially excited area. Details at higher magnifications of left depicted in 2. Oblique arrow, blood vessel (permits localization of areas photographed in 2). Arrowheads, normal ganglion‐type cells of mesencephalic trigeminal nucleus, surviving within completely depopulated region (for details on selective and absolute resistance of these ganglion‐type cells, see refs. 154,177). Lesion affected almost entire MRF, leaving intact only a limited area dorsolateral to RN. Lesion encroached on lateral part of CG and lateral part of deep layers of SC. With exception of ganglion‐type trigeminal cells, there was total loss of neurons within lesioned area. Sections stained with unsuppressed Nauta method showed integrity of axons in fields with complete neuronal depopulation. FTC, central tegmental field; III, oculomotor nucleus.

Adapted from Kitsikis and Steriade 396 and Steriade 867


Figure 14.

Firing rates and patterns of MRF neurons in cat. A: discharge rate as function of dorsoventral localization. 1: Location at frontal plane 2.5 of neurons whose firing rates were roughly estimated during recording and divided into 3 classes: silent, <3/s, and >3/s. AQ, aqueduct; 3, oculomotor n.; L, raphe linearis. 2: Dorsoventral position of 44 MRF neurons whose firing rates during quiet waking were measured in detail; Spearman rank correlation between high discharge rate and ventral location (ra = 0.42) significant at P < 0.005; arrow, limit between deep layers of SC and MRF. B: discharge patterns of rostrally projecting MRF cell, antidromically identified from zona incerta and synaptically driven from preoptic area and paramedian pons. Note sustained regular firing regardless of movements in waking (W+) or quiet wakefulness (W‐). ISIH, interspike interval histogram; N, number of intervals; X, mean interval; M, modal interval; C, variation coefficient; E, proportion of intervals in excess of depicted time range. Note symmetric shape and interval density near mode. Bottom: autocorrelations (ACFs) of same neuron during states of W‐, slow‐wave sleep (S), and desynchronized sleep (D). Note, during W‐, rhythmic firing with peaks at multiples of mode, and flat contours in S and D states.

Adapted from Steriade et al. 759


Figure 15.

Midbrain reticular formation neurons increase discharge rates in advance of behavioral and EEG signs of activated states in cat. A: electrographic criteria of transitional state (SW) from slow‐wave sleep (S) to wakefulness (W). Abrupt (1) and progressive transition, with an intermediate SW period (2). B: percent cumulative histograms (1‐s bins) of 2 MRF neurons (neuron 2 antidromically identified from CM‐PF). Abscissas, real time of recording; arrows and vertical lines, earliest signs of reduced amplitude and/or increased frequency of EEG waves (as in A, pt. 2, arrow indicates time 0 of SW period). Inflection points are seen to occur 10–22 s in advance of any change in EEG; overt signs of wakefulness (eye movements and increased muscular tone) appeared several seconds after arrows (as in A, pt. 2). C: increase in firing rate of MRF neurons of cat before end of S epochs developing into W. Left: percent cumulative histogram of neuron belonging to sample analyzed in graph depicted on right; arrows, first change in fully synchronized EEG waves. Right: 25 cells whose global mean rate in S was at least 4/s were analyzed during last minute of S in 42 epochs leading to SW or directly to W. Mann‐Whitney test was used to compare reference rate during first 30 s for all cells with their respective rates in the last 6 5‐s bins. Note significantly increased rates in the 3 5‐s bins before end of S (arrow) compared with discharge rate in the first 30 s.

Adapted from Steriade et al. 759,761


Figure 16.

Midbrain reticular formation neurons decrease discharge rates in advance of first electro‐graphic signs during transition from W to S. A: electrographic criteria of transitional state from W to S (WS). Graph shows median of firing rates in sample of 52 MRF cells during W, WS, and S; note that major and significant decrease in firing rate occurs from W to WS. B: neuron antidromically identified from CM‐PF. Top 2 traces, original spikes and EEG waves simultaneously displayed on oscilloscope. Note decreased firing rate, leading to neuronal silence, before first EEG spindle sequence (between arrows) during transition from W to S. Bottom traces, same activities during repeated EEG desynchronization‐synchronization transitions (polygraphic recordings). C: 12 transitions of unit firing with respect to start of EEG spindle (time 0). Arrow, level of discharge (median rate) during W. Asterisks, bins with significant (<0.05) decrease in firing rate compared with median rate during W. Significantly decreased firing rate occurred 1 s before spindle onset.

Adapted from Steriade 737,740 and Steriade et al. 759


Figure 17.

Cortical and thalamic neuronal activity during waking‐sleep states. A: unit discharges from cell in motor cortex of monkey during W and S. Surface EEG rhythms from motor and visual cortices shown below discharges. Note that cell remained active during S. B: patterns of firing of lateral geniculate cell in cat during W and S. Upper beam swept from below upward; sweeps triggered by spikes. Positive deflections upward for continuous beam, to left for swept beam. Different time calibrations for upper and lower beam. Note bursting discharges during S. C: discharge frequency in antidromically identified pyramidal tract (PT) neurons of monkey and in neurons that were not antidromically identified (non‐PT) during W, S, and D. Ordinates, spikes/s in PT and non‐PT cells. The PT and non‐PT cells are quite different with respect to both total amount of activity and changes in amount of activity as a function of sleep and waking.

A adapted from Jasper 341; B adapted from Hubel 324; C adapted from Evarts 209


Figure 18.

Discharge patterns of thalamocortical and corticothalamic neurons during waking‐sleep states in cat. A and B: intralaminar CL‐PC thalamic neurons antidromically identified from MC (neuron B was also synaptically activated from midbrain reticular core). Note high‐frequency spike clusters in S and their replacement by sustained discharge in both behavioral states of W and D. C: cortical cell recorded in area 5 and backfired from the CM thalamic n. (see inset at left with antidromic identification). Similarly sustained discharges in W and D; decreased firing rate in S. Bottom: polygraph traces depict unit spikes, focal EEG waves simultaneously recorded by microelectrode in area 5, EEG from depth of visual cortex, EMG of neck muscles, and EOG. Note tonically increased firing rate preceding onset of D (upward arrow) and during D, similar to discharge pattern in W (awakening marked by downward arrow). Two parts are separated by nondepicted period of 180 s. Tonic discharge of corticofugal neurons throughout D is very different from burst discharges of cortical interneurons related to REM epochs (see Fig. 20C).

A and B adapted from Glenn and Steriade 252; C adapted from Steriade 736


Figure 19.

Fast‐conducting and slow‐conducting PT neurons on arousal and subsequent steady waking in monkey and cat. A: histogram of antidromic response latencies of 67 precentral PT neurons after peduncular stimulation in monkey. Hatched columns, units in which spontaneous discharges stopped on arousal from synchronized sleep; white columns, neurons whose spontaneous firing increased on arousal. Inset: PT cells (a, b) that are represented in B. Arrowhead, stimulus; oblique arrow, lack of antidromic spike of cell b due to collision with spontaneous (S) discharge. B: 4 ink‐written traces depict a and b units, EEG waves, and eye movements. Note that fast‐conducting a neuron diminished firing rate on arousal but, afterward, during steady waking, increased its firing rate over value seen in synchronized sleep. Slow‐conducting b cell increased firing rate from beginning of arousal. C: intracellular recording of fast PT cell in midpontine pretrigeminal cat. Traces, from top to bottom: stimulation of MRF at 100/s, EEG from MC, hippocampal activity, zero potential level (broken line), cell activity (spike peaks off), spike rate/s, and cell activity at higher gain (spike peaks off). Bottom: antidromic responses to peduncular stimulation and spontaneous action potentials.

A and B adapted from Steriade et al. 747; C adapted from Inubushi et al. 329


Figure 20.

Discharge patterns of cortical interneurons during waking‐sleep states in cat. A: discharges during W, S, and D of putative short‐axon cell recorded from parietal association area 5, synaptically driven with high‐frequency spike barrage from LP thalamic n. (similar to interneuron depicted in C). B: similar type of cell in area 5. Polygraph traces, from top to bottom: unit discharges, EEG focal waves recorded by microelectrode, surface EEG waves, EMG, and ocular movements. Note diminution of spontaneous firing during arousal elicited by auditory stimulation (between arrows) and progressively increased occurrence of spike bursts with transition from W to S. C: another interneuron in area 5, driven by stimulation of LP thalamic n. with spike barrage at ∼500/s occurring in relation with onset of slow positive focal wave. Note stereotyped spontaneous spike bursts in all (W, S, and D) states. Note in polygraph traces (same activities as in B) that, during D (beginning at arrow), interneuronal discharge bursts are closely related to REM episodes and are dissimilar to sustained firing of corticofugal neurons during D, as shown in Fig. 18C.

Adapted from Steriade 736 and Steriade et al. 758


Figure 21.

Antidromic responsiveness of thalamocortical cells during waking‐sleep states in cat. A: inhibition during EEG spindles of antidromic responses in thalamocortical VL neurons. Two simultaneously recorded cells. Neuron b was superimposed on negative hump of neuron a; note that neuron b was antidromically activated in absence of discharge of neuron a (5). Figures on EEG record (1–6) indicate application of testing shock trains, corresponding to those in traces depicting evoked discharges. Abolition of antidromic responses occurs during EEG spindles (3 and 6). Evoked discharge of cell a disappeared 500 ms prior to EEG spindle (5). B, top: antidromic responses to area 5 stimulation in typical CL cell of this sample during W, S, and D. Arrowheads, stimuli. Bottom: median (arrowheads), mean (columns), and standard deviation for percentage of antidromic responsiveness (R) in sample of 19 CL‐PC intralaminar thalamic neurons backfired from pericruciate (areas 6 and 4) or anterior suprasylvian (area 5) gyri during W and S. Five of these neurons were also recorded during D. Asterisk, collision with spontaneous discharge.

A adapted from Steriade et al. 741; B adapted from Glenn and Steriade 252


Figure 22.

Antidromic responsiveness of corticofugal cells during waking‐sleep states in cat and monkey. A: facilitation of antidromic invasion of PT cell of cat during EEG desynchronization. Full responsiveness to a 4‐shock train during control period of spontaneous EEG desynchronization (1), depressed responsiveness during progressively developing EEG synchronization (2 and 3), and recovery of antidromic responsiveness during EEG desynchronization elicited by brief conditioning pulse train to MRF (4). B, top: corticothalamic cell in area 5 of cat, antidromically activated from n. CM. Antidromic responsiveness was investigated with a 4‐shock train. Arrowheads, stimuli. Bottom: percentage responsiveness (R) to 1st, 2nd, and 4th shock during W, S, and D. Below each state is mean rate of spontaneous firing (during another waking‐sleep cycle). C: pattern of antidromic invasion of pre‐central PT neurons during behavioral S (left) and arousal (right) in monkey. Dotted line and arrow indicate arousal. Fast‐conducting neuron has 0.5‐ms antidromic response latency to peduncular stimulation. 1 and 2, Responses to trains of 5 shocks (350/s) and 3 shocks (110/s), respectively, during 2 passages from sleep to arousal. Third trace: EEG desynchronization. Bottom trace: eye movements. Note spike fragmentation of first antidromic spike (arrowheads) during sleep and full recovery on arousal as well as diminution of spike fragmentation of successive responses.

A adapted from Steriade 735; B adapted from Steriade et al. 754; C adapted from Steriade et al. 747


Figure 23.

Primary synaptic excitation in thalamus and neocortex during waking‐sleep states in cat. Evoked potential studies (A‐C) and extracellular unit recordings (D‐F). Filled circles, testing shocks. A: simultaneous recording of field potentials evoked in lateral geniculate n. (LG) and at surface of visual cortex (VC) by optic tract stimulations (monopolar recordings). Response of LG consists of presynaptic [tract (t)] positive component and monosynaptically relayed (r) negative component. Different components of VC response numbered 1–5. Note, during 300/s stimulation of MRF, enhancement of monosynaptic (r) component in LG without alteration in presynaptic (t) deflection; note also increased amplitude of VC response. B: surface VC potentials evoked by stimulation of white matter beneath VC (1) or to deep layers in VC (2). Note that, with white matter stimulation, postsynaptic components of VC response are enhanced without changes in presynaptic component. C: simultaneous recording of field potentials evoked in the VL thalamic n. and MC by stimulation of cerebellothalamic pathway during W, S, and D. Note in S, compared with both W and D, obliteration of monosynaptic thalamic wave (r) and marked reduction of cortical response, without alteration of component t that reflects activity in afferent fibers to VL n. (in this case, bipolar recording in VL). D: MRF‐evoked synaptic discharges in intralaminar CL thalamic neuron antidromically identified from parietal association area 5. Note decreased latency, increased probability of discharges in early bins, and shorter duration of spike bursts during W compared with S. E and F: unit recordings in SI cortex (SC) and white matter stimulation in animals with complete lesion of VB thalamic complex. The MRF shock train preceded testing shocks by 3–5 ms (not depicted). Note MRF‐induced facilitation of evoked discharges. F: MRF potentiation is seen by decrease in latency of discharges to first stimulus and increased discharge probability to second stimulus.

A and B adapted from Steriade 733,734; C adapted from Steriade et al. 752; D adapted from Steriade and Glenn 750; E and F adapted from Steriade and Morin 756


Figure 24.

Effects of natural arousal and midbrain reticular stimulation on inhibitory processes in thalamus and neocortex of cat and monkey. A: periods of suppressed firing (a, b, and c) and postinhibitory rebound excitation of neuron recorded in rostrodorsal part of LP thalamic n. after single‐shock stimulation (arrowhead) of cortical area 5 in cat. Preceding shock train to MRF reduced duration of periods a and b and abolished silent period c. B: patterns of VL‐evoked events in precentral neuron of monkey during S and W. Note that first inhibitory period evoked by VL stimulus (dots) persists in W, but subsequent rhythmic inhibition‐rebound sequences seen during S are abolished in W on background of increased spontaneous discharge. C: method of testing recurrent inhibition acting on antidromic discharges elicited in cat PT neurons by pes peduncular (PP) stimulation. Midbrain lesion included medial lemniscus to avoid stimulation of afferent fibers. Conditioning volley (C) was delivered at 13 V [minimal voltage required to elicit inhibitory effects on testing (T) response induced by shock at 5 V, which is minimal voltage required to evoke 100% antidromic invasion]. At paired C‐T stimulation, complete inhibition of T response or spike fragmentation (arrow). Graph depicts much longer inhibition with 3 PP shocks than with single shock. With both conditioning procedures (1 PP and 3 PP), recovery was slower during S than during W. D: inhibition of synaptic discharges evoked by stimulation of posterior part of VL in precentral PT neuron of monkey. Left: field positive (inhibitory) wave evoked by first VL stimulus and facilitation (during W) of evoked discharges by second stimulus at 75‐ms interval toward end of inhibition. Right: percentage responsiveness of discharge evoked by first stimulus (time 0) and by second stimulus at 3 time intervals (15 ms, 27 ms, and 75 ms) during W and S.

A adapted from Steriade et al. 757; B adapted from Steriade et al. 748; C and D adapted from Steriade and Deschěnes 745


Figure 25.

Secondary excitation and incremental responses during waking‐sleep states in cat. A: focal slow waves recorded at depth of 0.5 mm in cortical area 5 after stimulation of rostrodorsal part of LP thalamic n. during W, S, and D (50 averaged sweeps). Note during S selective enhancement of second (b) depth‐negative component. B: responses of cortical SI neuron to VB thalamic stimulation during S and W. Note during W increased probability of VB‐evoked early discharge and suppression of late repetitive discharges. C: simultaneous recording of field potentials (1: 50 averaged sweeps) and unit discharges (2) at depth of 1 mm in suprasylvian area 5, evoked by 2 0.1‐s‐delayed stimuli to LP thalamic n.; stimuli indicated by dots in 1 and by vertical bars in 2. Secondary depth‐negative wave b associated with repetitive discharges is selectively enhanced at second stimulus. D: poststimulus histograms of augmenting responses in sample of 7 cortical SI neurons, elicited by white matter (WM) stimuli at frequency of 10/s in VB‐lesioned preparation. S, stimulus number; R, responsiveness (total number of discharge to 100 shocks). Note that augmentation elicited by S‐2 (the second 0.1‐sdelayed shock) consists of increased secondary excitation (10–15 ms, arrow) simultaneous to decreased probability of primary synaptic excitation. Effect of a conditioning stimulation of MRF consists of an increased probability of primary excitation and a change in pattern of augmenting response to second stimulus into a primary one. E: augmenting field potentials at depth of 0.7 mm in area 5 to 10/s stimulation of LP thalamic nucleus during W, S, and D. Arrowheads in A, B, C, and E indicate stimuli.

A and E adapted from Steriade 738; B adapted from Steriade 735; C adapted from Steriade 736; D adapted from Steriade and Morin 756


Figure 26.

Cortically evoked rhythmic hyperpolarization‐rebound sequences. Intracellular recording of VL thalamic relay cell, antidromically invaded from precruciate MC and monosynaptically driven from brachium conjunctivum (BC). Resting membrane potential: −55 mV. 1: Typical response sequence. 2 to 4: 40 Averaged sweeps. Note in 3 and 4, analyzed at same speed, powerful rhythmic activity within frequency of spindle waves (∼7.5/s) evoked by MC, contrasting with lack of such activity after BC stimulation. Arrowheads, MC stimulation; filled circle, BC stimulation.

Adapted from Steriade 740


Figure 27.

Effect of membrane potential on excitability of VL thalamic relay neuron in cat; intracellular recording. A and B: antidromic and orthodromic activation by MC and BC stimulation, respectively. C: direct stimulation of cell by current pulse. At resting potential (0 nA) current pulse was subthreshold for spike initiation. During injection of 1 nA continuous hyperpolarizing current (2), same pulse remained subthreshold but break was followed by slow decaying response on voltage trace. Finally, under 2 nA hyperpolarizing current (3), pulse triggered “slow spike” and burst of action potentials. D: postanodal exaltation phenomenon obtained by passing hyperpolarizing current pulses of increasing intensities at resting potential. Note stereotyped burst response in D, pt. 3 compared with C, pt. 3. E: traces, same phenomenon as in D but at higher gain and slower speed. Current intensities similar to those in D. E, pt. 2: slow spike can be seen in isolation. Upper calibration bars apply to A and B.

Adapted from Deschěnes et al. 182


Figure 28.

Effects of rostral reticular stimulation on intracellulary recorded inhibitory processes in thalamic neurons. A: excitatory postsynaptic potential (EPSP) and inhibitory PSP (IPSP) sequences in VL neuron of cat during 7/s midline thalamic stimulation (1) and marked attenuation of hyperpolarizing potentials and increase in cell discharges during simultaneous low‐frequency thalamic stimulation and high‐frequency brain stem reticular stimulation (2). Upper traces of 1 and 2, evoked responses at motor cortical surface. B: VL relay cell in cat. 1: MC stimulation (arrowheads) evokes antidromic discharge followed by rhythmic inhibitory‐rebound sequences. 2: Conditioning 320/s shock train to MRF leaves intact the cortically elicited early IPSP but abolishes late phase of hyperpolarization and following rebound‐inhibition sequence. 3: Effect of MRF stimulation alone. C: slow (∼0.1 Hz) thalamic rhythm of hyperpolarizing episodes and its suppression during MRF stimulation. 1: Ink‐written intracellular recordings of VL relay cell. Note oscillations within frequency range of spindles (∼7 Hz), consisting of phasic hyperpolarizations followed by rebounds, appearing during slow rhythm (−0.12 Hz) of hyperpolarizing episodes, each lasting 2.3 s. Postinhibitory rebounds consist of slow spike (see Fig. 27) superimposed by fast repetitive action potentials. 2: Suppression of slow rhythm of thalamic hyperpolarizations during MRF stimulation (brief shock train every second); note increased background firing that outlasted MRF stimulation period; first episode of hyperpolarization after MRF stimulation lasted only 1.3 s, compared with 2.3‐s duration of these episodes before MRF stimulation.

A adapted from Purpura et al. 612; B and C adapted from Steriade 740


Figure 29.

Activities of RE neurons during waking‐sleep states in cat. Two neurons recorded in rostral pole of nucleus reticularis thalami, monosynaptically driven from precruciate gyrus with high‐frequency spike barrages, as shown for neuron A in top poststimulus histogram. T, number of trials; X, mean latency (ms); M, latency mode (ms); C, coefficient of variation; R, responsiveness (total number of spikes to 100 stimuli during depicted time). Traces in A: unit activity, surface cortical EEG waves, focal EEG waves recorded from electrode pair that induced synaptic activation of the neuron, EMG, and time (1 s). Note spike bursts closely related to EEG spindles and tonically increased firing rate during arousal and sustained waking. B: peristimulus histograms depict activities evoked from precruciate cortex during S and W. Left histograms (2‐ms bins) show details of early excitation; right histograms (25‐ms bins) show late inhibition‐rebound phases. Note in left histograms that probability of early evoked discharges (first 2‐ms bin) doubles in W compared with S. In S evoked burst mostly extends within latency range of secondary excitation (∼15–40 ms). Also note (right histograms) 2 evoked rebound sequences in S, whereas such events are lacking in W (see similar phenomena in Figs. 24A and 28B for thalamic relay cells).

Adapted from Steriade 740


Figure 30.

Facilitation of monkey's parietal visual neurons by attentive fixation. A: comparison of responses of parietal light‐sensitive neurons to visual stimuli in no‐task and task modes. Absence of responses during no‐task mode compared with strong responses in task mode is obvious. Summing histograms: standard error of mean calculated for each bin of the histograms; value shown by dotted line. Corresponding bin pairs within the histograms tested for significant differences (t test); bin pairs marked with diamonds differed at 5% level of significance. Overall response in no‐task and task states compared in following way. Rate of impulse discharge in prestimulus period was subtracted from that in poststimulus period for each trial, and populations of remainders were tested for significant differences (P ≪K 0.05 required) and used to form facilitation ratio for each neuron. Ratios for neurons with significant differences plotted in B. Facilitation ratio is that between net increment in response evoked by light stimulus in state of interested fixation over that evoked by physically identical and retinotopically similar stimulus delivered in no‐task or intertrial states. Fifty‐one neurons showed ratios of ≥1.0, and of these, difference was significant at 5% level (t test) for 38; values plotted in histogram. Ratio was fractional for 4 neurons indicated at left: for them, response was significantly greater in no‐task state than during interested fixation.

Adapted from Mountcastle et al. 528


Figure 31.

Facilitatory action of acetylcholine (ACh) in cerebral cortex. A: intracellular record from neuron in motor area of cat shows delayed depolarizing effect and prolonged firing evoked by iontophoretic application of ACh (140 nA). B: brief and instantly reversible depolarization and strong firing of same neuron caused by short intracellular current injection (monitored on lower trace). C: same neuron depressed after treatment with dinitrophenol; no longer fired in response to application of ACh (as in A); however, during continued ACh application, identical intracellular current injection (cf. B) now induced particularly powerful and prolonged discharge. The ACh thus greatly facilitates and prolongs any depolarizing input received by same cell. D and E: magnitude and time course of changes in potential and resistance induced by ACh. Open circles, resting potential: note slow and prolonged depolarizing effect; open triangles, resting resistance: note marked increase in resistance synchronous with depolarization; closed symbols, corresponding data recorded during IPSPs: they show relatively little change, except some possible reduction of inhibitory effect. F: summary of probable mechanism of action of ACh. In contrast to situation at neuromuscular junction (and other sites of nicotinic cholinergic action) where depolarization is mediated by enhancement of Na+ permeability, muscarinic ACh action in cerebral cortex tends to reduce membrane K+ permeability, causing an increase in overall resistance and facilitating depolarizing effect of other depolarizing inputs, such as EPSPs, which probably act by increasing Na+ permeability.

Adapted from Krnjević et al. 409


Figure 32.

Eye movements (EM), EEG, systolic blood pressure (SBP), respiration (resp), pulse, and body movements (BM) in a 100‐min sample of uninterrupted sleep over successive minutes of typical sleep cycle. Entire interval from minute 242 to 273 is considered to be REM period, even though eye movements (heavy bars) are not continuous.

Adapted from Snyder et al. 722


Figure 33.

PGO waves and relation to REM sleep and eye‐movement direction. A: NREM‐REM transition showing PGO waves in LGB (types I and II). During transition periods from NREM to REM sleep, biphasic (PGO) waves in LGB first appear as large single events (type I waves). Waves become clustered and of diminished and decrementing amplitude (type II waves) as signs of REM sleep become more prominent: atonia (EMG), desynchronization of cortical EEG (Cx), hippocampal‐θ (HIP), and REMs (EOG). B: side‐to‐side alternation of primary waves. Once REM period is well established, PGO wave amplitude primacy alternates from one geniculate to the other according to lateral direction of eye movements. When there is rightward movement of the eyes (EOG‐R), corresponding PGO wave cluster is larger in right LGB (dots) than in left. Conversely when there is leftward movement of the eyes (EOG‐L) waves are larger in left LGB (dots).

Adapted from Nelson et al. 856


Figure 34.

Behavioral state and neuronal discharge rate. Mean rates of brain cell populations shown as function of behavioral state. A: D‐on cells. Most cells of the brain have higher rates in D and W than in S. Cbm, cerebellar Purkinje cells; VN, vestibular n; Thal, thalamic n.; PRF, pontine reticular formation; Ctx, cerebral cortex; Hypo, hypothalamus. B: D‐off cells. Minority of cells that decrease rate in D are all found in or near aminergic zones of pontine brain stem. Ret N, reticular n. subpopulation; RN, raphe n.; PN, peribrachial neurons.

Adapted from Steriade and Hobson 751


Figure 35.

Relation of pontine burst cell firing, PGO wave amplitude, and eye movement lateral direction. A: burst cell and waves. Recording of cell in peribrachial region of pons that fires 4 clusters of spikes, each of which precedes PGO waves in LGB by 10–20 ms. Geniculate ipsilateral to cell (LGB1) shows markedly greater wave amplitude than the contralateral geniculate (LGBc) in initial pair of series. Correlation between pontine burst cell activity and LGB wave amplitude is 0.99. B: efferent copy. Three‐way correlation between eye movement direction (EOG), geniculate wave amplitude (LGB), and pontine burst cell activity shown on drawings of ventral brain surface (above) and exemplified in oscilloscope tracings (below). When eye movement is toward same side as cell, wave amplitude is greater in LGB1 and vice versa.

Adapted from Nelson et al. 856


Figure 36.

Pontine localization of REM sleep generator: lesion evidence. A: prepontine transection of the brain (with forebrain ablation) is followed by alternation of state resembling waking and REM sleep. Note suppression of muscle tone (EMG) and clustered eye movements (EOG). The EEG of brain stem shows low‐amplitude PGO waves occurring with eye‐movement clusters. Data show that REM sleep clock and trigger neurons are retropontine. B: in isolated pons preparation, C1 spinal cord transection is added to prepontine brain stem transection. Atonia is eliminated but REMs and pontine PGO waves appear with 30‐min periodicity characteristic of REM sleep in cat. Data show that REM sleep clock and trigger must reside in lower brain stem.

Adapted from Matsuzaki 474


Figure 37.

Pontine mechanisms of REM sleep generation: D‐on cells and executive actions. A: identification. Reticular neurons projecting to spinal cord can be identified by antidromic activation from axons at lumbar spinal cord levels via chronically implanted stimulating electrodes while recording somatic action potentials in brain stem of head‐restrained but unanesthetized cats with moveable microelectrode. At paramedian pontine site shown in 1, cell was recorded that showed a fixed‐latency response (2), fast following (3), and collision (4) with orthodromic spikes. Cell was typical in that it showed lowest firing rate in waking (5), a slightly greater rate in S (6), and intense bursting prior to REM bursts of D (7). B: selectivity. Tendency to concentrate firing in D quantified as ratio of mean rate in that state to rate in W or S. As shown in 1‐min episodes of spontaneous firing for each of these cells in the 3 sites, selectivity was highest in pontine reticular formation (1), next highest in pontomesencephalic reticular formation (2), and lowest in such precerebellar nuclei as tegmental reticular nucleus (TRC) (3). C: phasic latency. Tendency to fire prior to eye movements quantified by plotting sequential and cumulative histograms as shown and averaging across populations. Longest phase leads (up to 300 ms) and greatest increases (up to 100% of firing) were seen in pontine reticular formation. D: tonic latency. Pontine reticular neurons showed statistically significant increases in firing rate as early as 3.5 min prior to D epochs. These tonic rate changes were longer in gigantocellular tegmental field (FTG) than any other neuronal group in brain stem or forebrain. Note that rate increase occurs in episodic increments suggesting that phasic activation waves may spatially and temporally summate as recruitment spreads within the reticular pool. E: periodicity. Spontaneous discharge of single FTG neuron in 10 h of continuous recording. Discharge peaks occur at ∼30‐min intervals coincident with each REM period. F: proportion of firings by 4 single FTG neurons averaged in successive intervals of repeated sleep cycles. Each cycle was time normalized by establishing duration of D and dividing it into 5 equal parts; the 10 min before (D‐10) and after (D+10) each D period were also examined. For each cycle, percentage of total number of firings was determined for each of 25 bins.

Adapted from Wysinski et al. 833, Hobson et al. 308, Hobson et al. 309, Pivik et al. 595, and McCarley and Hobson 480


Figure 38.

Pontine mechanisms of REM sleep generation: D‐off cells and permissive action. In contrast to discharge pattern and activity profile on D‐on cells found in pontine reticular formation is behavior of neurons in aminergic nuclei such as LC cells. A: anatomical location. Histological reconstruction (drawing) and computer plot (inset) of microelectrode penetration and recording site, in n. LC (CAE), of D‐off cell. FTG, pontine reticular formation; 7G, genu, of facial nerve; SA, striae acousticae; VSL, lateral vestibular nucleus. B: spike trains. The LC cell recorded at site shown in A fired regularly in W, intermittently in S, and arrested discharge in D. Cell resumed firing, with no change in spike height, during subsequent cycle. C: dotgrams. One‐min periods of spontaneous activity in each of the 3 states shown by triggering Z‐axis of oscilloscope with action potentials shown in B. D: cell activity curves. As in D‐on cells, change in rate across states is progressive and continuous in D‐off cells but opposite in direction. Resumption of firing prior to end of D‐sleep epochs shown in insets of both graphs. SWS, slow‐wave sleep; PS, paradoxical sleep.

A, B, and C from J. A. Hobson, unpublished observations. D adapted from Aston‐Jones and Bloom 46


Figure 39.

A: reciprocal activity of LC and giant reticular neurons during sleep cycle. Reciprocal discharge by cells in CAE 559,560,561,562 and FTG 563,564,565,566,567. Outline tracing, sagittal section of cat pontine brain at 2.5 mm lateral to midline, showing path of an exploring microelectrode passing through cerebellum into dorsal brain stem. Inset, location of the 7 successive recording sites in penetration; circle diameter is 5 mm. Cumulative discharge histograms of cells recorded at 7 sites shown in inset during transition period beginning 2 min after D onset (vertical line) and ending 1 min thereafter. Each activity curve shows cumulative percentage of discharge for as much of the epoch as was free of arousal. Units recorded in LC and subcoeruleus decrease discharge rate (negatively inflected histograms), whereas those in FTG increase discharge rate (positively inflected histograms) with approach and onset of D. B: pooled geometric mean rate for population of 21 LC and 34 FTG neurons. Abscissa is W, S, and D. C: selectivity ratios of geometric mean rate for pooled population data shown in B. D: pooled cumulative histograms of discharge rates in subsets of the 2 neuronal groups during transition period from S to D. Ordinate, percentage of discharge; abscissa, time in minutes, where to is onset of D and 2 min before and 1 min after are shown. Overall mirroring of the 2 curves suggests that activity of the population may be modulated by shared, perhaps mutual, process.

Adapted from Hobson et al. 310


Figure 40.

Reciprocal interaction model of REM sleep generation. A: anatomical maps. Locus of D‐on cells (hatched area) in reticular formation nuclei and D‐off cells (dashed line) in pontine aminergic nuclei is shown on sagittal section of brain stem. Four sites with neurons relatively indifferent to state are shown on whole brain schematic: cerebral cortex (Ctx), cerebellar Purkinje cells (Cbm), other reticular nuclei (Ret), and PG. CAE, locus coeruleus; FTP, L, posterior and lateral tegmental fields; Pbl, parabrachial zone; RN, raphe n. B: selectivity gradient. The D/W selectivity ratios, computed and pooled as described in Fig. 39, are plotted for some of nuclei shown in A. Continuous gradient from most positively selective cells in FTG to most negatively selective cells in RN. Converse gradient would describe propensity to fire in waking under same recording conditions. C: physiological model. D‐on cells are assumed to be cholinergic/excitatory (E); D‐off cells are assumed to be aminergic/inhibitory (I). Each group is assumed to have both feedforward and feedback interconnections giving structural and synaptic model of reciprocal interaction. Model can be described mathematically and tested pharmacologically.



Figure 41.

Reciprocal interaction model: physiological and mathematical aspects. A: structural model of interaction between FTG and LC cell populations. Plus sign implies excitatory influences; minus sign implies inhibitory influences; a‐d correspond to constants associated with strength of connection and included in text equations. B: theoretical curve derived from model that best fits FTG unit in Fig. 37E. In this fit a = 0.3029, c = 0.1514, and the initial conditions (amplitude unsealed) were x(0) = 1 and y(0) = 4.5. C: histogram is average discharge level of another FTG unit over 12 sleep‐waking cycles, each normalized to constant duration. Cycle begins with end of D. Arrow, bin with most probable time of D onset. Solid curve, FTG fit; dotted curve, LC fit derived from model with values a 0.5490, c = 0.2745, x(0) = −1, and y(0) = 3.0 for the 2 populations. Dot, equilibrium value for the 2 populations. D: geometric mean values of discharge activity of 10 LC cells before (S), during (D), and after (W) a D episode. Each time epoch is equal to one‐quarter of D period. Note that discharge‐rate increase begins in last quarter of D, at same time that FTG activity is falling as predicted by model.



Figure 42.

Modulation of pontine neuronal excitability over sleep‐waking cycle. The finding that pontine reticular neurons fire in association with movement during waking in unrestrained cats is shown to be consequence of strong excitatory drive from other central motoneurons and afferent input from periphery. That different mechanisms could lead to same net output in W and D may be appreciated by comparing synaptic models (A) with activity pattern (B). In W powerful afferent excitation of FTG (++++) overcome tonic inhibitory restraint coming from aminergic neurons (—) and summates with recurrent collateral excitation (++) to produce clustered firing. In absence of afferent drive, cell is silent in W. In D less powerful afferent input (++) is capable of driving cells because level of inhibitory restraint has decreased (−). For same reason collateral excitation is more capable of sustaining firing within reticular pool. Result is repeated clustered firing that drives eye movements and muscle twitches. Because of tonic inhibition, motoneurons of large skeletal muscles do not respond; hence there is no major movement and no feedback from movement to central motor pattern generators in D. Because of phasic presynaptic inhibition of la cutaneous afferents there is also no change from peripheral sensory stimuli to drive system as is present in W. In slow‐wave sleep the system is in condition intermediate between the 2 extremes just discussed: disinhibition increases progressively and facilitation is markedly reduced so that cell firing is sparse.



Figure 43.

Pharmacology of REM sleep generation: comparison of results with predictions of reciprocal interaction model. REM sleep is favored (black postsynaptic neurons) by either enhancement of cholinergic transmission or by blockade of noradrenergic and serotonergic neurotransmission. Top: REM sleep increases are seen when neurotransmission is directly enhanced by agonists such as carbachol, which binds and stimulates acetylcholine receptors; REM sleep is also enhanced by eserine, which prevents enzymatic breakdown of endogenously released acetylcholine by cholinesterase. In contrast, REM sleep decreases are seen when muscarinic acetylcholine receptor is competitively occupied by atropine. Middle: REM sleep is decreased by any drug action that increases effective amount of this neurotransmitter. Monoamine oxidase (MAO) is inhibited (MAOI), leading to increase in NE and, via an inhibition of REM generators, a decrease in REM. In contrast, when β‐receptors are blocked by propranolol, postsynaptic REM generator cells are disinhibited and REM sleep is increased. Bottom: reuptake of aminergic neurotransmitter (e.g., by amitriptyline) may also lead to effective increase in serotonergic inhibition of REM sleep. On other hand, when reserpine depletes amount of serotonin (and NE) in vesicles, there is less inhibitory restraint and REM sleep increases.

Adapted from Vivaldi et al. 805


Figure 44.

Cholinergic REM sleep: enhancement effects of carbachol. A: predictions of reciprocal interaction model. Under physiological conditions, REM sleep occurs when D‐on cells are exposed to high levels of cholinergic excitation and low levels of aminergic inhibition. Carbachol increases cholinergic excitation of D‐on cells and thereby tips balance of system toward REM sleep generation. B: polygraph records. Qualitative identity, in all channels of physiological and pharmacological D. In both cases, EMG is flat, indicating atonia; cortical EEG (CX) is desynchronized, indicating activation; there are PGO waves in EEG of LGB; there is θ‐activity in hippocampal EEG (HIP); and REMs are indicated by deflections of EOG. C: individual time course data. Compared with control (above) note onset (<10 min) of prolonged (>60 min) D episodes seen after each of 3 carbachol cannula placements in pontine reticular formation. D: pooled time course data. Consistent enhancement of D seen across carbachol trials and subjects. At 1 h, peak value of 75% is 4–5 times greater than control and the effect lasts for 3 h.



Figure 45.

Cholinergic REM sleep enhancement: ionotophoretic effects of carbachol. A: chemical microstimulation. Microiontophoretic drug delivery system. Head restraint implant placed stereotaxically so that glass microiontophoretic pipette can be directed at pontine brain stem under chronic recording conditions. Electrographic D with carbachol (D‐carb) can be produced with currents as low as 300 nA for 10 min, and action potentials of single giant cells can be recorded through same pipette before and after delivery of drug. B: microiontophoretically induced D‐carb. Electrographic signs after delivery of carbachol by passing current through glass micropipette indistinguishable from those of physiological D and from D‐carb induced in cannula diffusion experiments. Injection was histologically localized to rostral pole of paramedian pontine reticular formation. C: iontophoretically induced PGO waves. At site different from that of B, note stereotyped groups of 4–8 waves that begin to occur, in absence of other D signs, at 33–52 min and are still present at 13 h, 42 min postinjection. Data suggest that it may be possible to chemically microdissect neuronal ensembles responsible for D‐state components. D: cross‐correlation of iontophoretically induced PGO waves and cell activity. Average PGO waveform and activity of single cell cross‐correlated in tape‐recorded segments of data shown in C. Note identity of PGO waveform (upper trace) to that seen in physiological D. Lower trace: histogram in which peak of PGO waves is taken as time zero and frequency of unit firing is counted for 500 ms before and after wave peak. Unit activity shows sharp increase in 40 ms prior to wave peak. This neuron histologically localized to PGO burst cell zone (see Fig. 35A).

Adapted from Vivaldi et al. 805


Figure 46.

Cholinergic REM sleep enhancement: intrabrain differentiation of enhancement and suppression stimulation sites. A: enhancement sites. Left: anterodorsal pontine reticular formation from which pure REM sleep is evoked. Middle: posteroventral pontine reticular formation from which REM sleep with stereotyped side‐to‐side alternation of eye movements (EMs) and PGO waves are evoked. Right: peribrachial pontine tegmentum from which state‐independent PGO waves are evoked. B: suppression sites. Left: midbrain reticular formation from which arousal with persistent motor circling is evoked. Middle: medullary reticular formation from which arousal with axial trunk rolling is evoked. Right: periabducens pontine tegmentum from which arousal with persistent oscillating eye movements (OEMs) is evoked. In each diagram parasagittal distance (in mm) is indicated. For sections at 1.2 mm: MLB, medial longitudinal bundle; TV, ventral tegmental n.; 6N, abducens nerve; TG, genu of facial nerve; 6, abducens nucleus; PH, praepositus hypoglossi. For section at 3.7: FTP, paralemniscal tegmental field; 5M, motor division of trigeminal nucleus; 7N, facial nerve; SO, superior olive; 7, facial nucleus; SA, striae acousticae; VL, lateral vestibular n.; S, solitary tract; VS, superior vestibular n.; EMG, nuchal EMG; EEG, cortical EEG; LGB, lateral geniculate body EEG; EOG, interorbital EOG. Each record ∼20 s.

H. Baghdoyan and J. A. Hobson, unpublished observations


Figure 47.

Cholinergic REM sleep enhancement: effects of bethanechol. A: molecular configuration. Top: acetylcholine, naturally occurring neurotransmitter that is rapidly degraded by cholinesterase. Middle: carbamylcholine (carbachol), long‐acting synthetic agent that resists enzymatic degradation. Carbachol is mixed nicotinic and muscarinic agonist. Bottom: β‐methyl carbamylcholine (bethanechol), long‐acting synthetic agent but pure muscarinic agonist. B: polygraphic records. EOG, flurries of REMs; EMG, complete suppression of nuchal muscle tone; EEG, desynchronization of cortical electrical activity; LGB, biphasic waves in LGB. Bethanechol‐induced state appears to be an intensified version of naturally occurring D state. C: time course data. Top: bethanechol trials and matched saline control recordings for 3 cats. At latencies of 10–30 min, D periods of up to 55 min ensue and are virtually continuous for 2 h. In all cases, latencies are shorter, durations longer, and intervals shorter than control values. Bottom: D‐beth values shown in top were pooled with 30‐min bins. Peak (85%) was reached at 1 h followed by progressive decline to trough (30%) at 3.5 h but with subsequent peaking. First time at which control levels matched experimental values was 5.5 h postinjection. D: anatomical differences indicate site specificity of enhancement and suppression of D by bethanechol. Enhancement of D was seen only at pontine sites where 4 of 7 trials produced at least a doubling of saline control values and only 1 produced suppression. In contrast, suppression was rule at medullary and midbrain sites where 6 of 8 trials produced at least a halving of control values and none produced enhancement. Four of the 6 medullary trials produced complete suppression of D.

M. Goldberg, E. Vivaldi, D. Riew, and J. A. Hobson, unpublished observations


Figure 48.

Antiaminergic REM‐sleep enhancement: effects of propranolol. A: predictions of reciprocal interaction model. Under physiological conditions, noradrenergic inhibition restrains cholinergic D generator. With β‐adrenergic blockade by microinjection or propranolol, D generator is disinhibited. B: polygraphic records show natural D onset and that seen after propranolol. There is no synchronized electroencephalographic activity in the cortex (CX) so that drug‐treated animal enters D directly from W. Channels as in Fig. 44. C: individual time course data. Enhancement of D by propranolol occurs by way of more episodes of normal length indicating that either threshold is not as effectively lowered as with carbachol and/or that serotonergic inhibition is capable of ending drug‐induced episodes. D: pooled time course data. Peak effect, at 1 h, is 2–3 times control. After return to base line at 3 h, there appears to be an undershoot or suppression that is not seen with cholinergic agonist enhancement.

Adapted from Vivaldi et al. 805


Figure 49.

Pathophysiology of narcolepsy. Narcoleptic patients have several abnormalities of sleep‐waking state control that appear to be result of changes in set point of REM state oscillator. During waking, REM sleep signs intrude as sleepiness, sleep attacks, and cataplexy. This increased propensity to REM is measured as decreases in multiple‐sleep latency test and may manifest itself in hypnogogic hallucinations or by sustained sleep‐onset REM period (normal subjects go quickly through stage I). On arousal from REM sleep, there may be corresponding persistence of mental phenomena of REM (hypnopompic hallucinations) and/or REM sleep motor inhibition (sleep paralysis). Reciprocal interaction model accounts for all these phenomena by hypothesizing decreased level of aminergic inhibition (light dashed line) and/or increased level of cholinergic excitation (light solid line) in pontine oscillator. During waking there is an increased propensity for REM generator to reach threshold. At sleep onset there is brief escape from aminergic restraint and sleep‐onset REM period is triggered. System then resets and cycles normally until end of REM when there is a lag in reinstitution of waking‐state conditions and REM phenomena again escape their normal temporal bounds. Clinical efficacy of aminergic agonists (e.g., amphetamine) or amine reuptake blockers (e.g., imipramine) may be due to their capacity to reset aminergic inhibition to normal level (heavy dashed line). By reciprocal interaction, this would also reset level of cholinergic generator (heavy solid line).



Figure 50.

Pathophysiology of sleep apnea syndrome. During waking, respiratory oscillator of medulla receives tonic drive from other neural structures and can respond to voluntary and metabolic signals to change breathing pattern. Ventilation is assured by active maintenance of airway pathway via tonus of oropharyngeal musculature. In NREM sleep, central drive on both respiratory oscillator and peripheral muscles declines due to disfacilitation. Respiratory rate and amplitude thus diminish and airway is subject to collapse. If obstruction occurs, forced expiratory effort may actually aggravate airway construction and prolonged apneas with marked hypoventilation and hypoxia may occur. During REM sleep, activation of pontine generator neurons produces tonic and phasic driving of respiratory oscillator, which may desynchronize leading to hyperpnea and/or to apnea. In addition, medullary oscillator becomes unresponsive to metabolic signals. In patients with tendency to airway collapse, these processes may multiply deleterious effects of ventilation.



Figure 51.

Motor activity in sleep: timing of posture shifts. A: posture shifts in video data. Movements are scored by making frame‐by‐frame comparisons of posture on replay of tape. Left, top and bottom: unambiguous posture shifts (>90% trunk rotation). Right, top and bottom: no such postural adjustments. Right, top: no movement is discernible; bottom: arm but no trunk movement. Drawings were made by tracing outlines of human subject in time‐lapse photographic study. B: individual records. Movement profiles characteristic of prompt (top) and delayed (bottom) sleep onset. Visual cross‐correlation between EEG data and posture shifts was established for each subject‐night. Movement ceases when NREM stages are progressive (top) but persists when waking and stage I alternate (bottom). Stereotyped coordination of posture shifts and sleep‐stage sequence followed sleep onset. Descending stages of NREM sleep, especially those reaching stages III and IV, were movement free. However, ascending NREM stages and REM (black areas) interruptions or terminations were accompanied by movements. C: pooled movement data. Immobility and sleep cycle phase. Top: average of 44 sleep cycles. Bottom: time of beginning (left curve) and ending (right curve) of 44 epochs of postural immobility related to average of all sleep cycles in which they occurred (top). Each curve is cumulative histogram of percentage of occurrences of immobility as function of percentage of cycle completed. Note steep and smoothly ascending curve of onsets indicating that immobility begins in association with early NREM sleep; curve of endings is by contrast inflected sharply at stage IV onset, which indicates that process controlling posture shifts is activated well before end of NREM sleep. D: neuronal activity curve. Actual activity of presumed REM generator neuron in cat brain stem (solid stepped curve) is compared with theoretical curves of on‐cell population (solid curve) and off‐cell population (dotted curve) as function of percent of cycle completed. Comparing D with C reveals inverse parallelism between timing curve of onsets in C of immobility and off‐cell trajectory in D; it is hypothesized that motor disfacilitation is occurring during first third of cycle. Later, curve of brain stem neuronal activation (D) directly parallels immobility offset in C. Both appear to be manifestations of phase shifts in motor systems such that immobility of REM sleep is produced by active peripheral inhibition in face of strong central facilitation.

Adapted from Aaronson et al. 1


Figure 52.

Mental activity in sleep: psychophysiology of dreaming. A: systems model. As a result of disinhibition caused by cessation of aminergic neuronal firing, brain stem reticular systems autoactivate. Their outputs have effects including depolarization of afferent terminals causing phasic presynaptic inhibition and blockade of external stimuli, especially during bursts of REM, and postsynaptic hyperpolarization causing tonic inhibition of motoneurons that effectively counteract concomitant motor commands so that somatic movement is blocked. Only oculomotor commands are read out as eye movements because motoneurons are not inhibited. Forebrain, activated by reticular formation and also aminergically disinhibited, receives efferent copy or corollary discharge information about somatic motor and oculomotor commands from which it may synthesize such internally generated perceptions as visual imagery and sensation of movement, both of which typify dream mentation. Forebrain may in turn generate its own motor commands, which help to perpetuate process via positive feedback to reticular formation. B: synaptic model. Some directly and indirectly disinhibited neuronal systems, together with their supposed contributions to REM sleep phenomena. At level of brain stem, 4 neuronal types are illustrated. MRF: midbrain reticular neurons projecting to thalamus that convey tonic and phasic activating signals rostrally; PGO: burst cells in peribrachial region that convey phasic activation and specific eye movement information to geniculate body and cortex (dashed line indicates uncertainty of direct projection); PRF: pontine reticular formation neurons that transmit phasic activation signals to oculomotor neurons (VI) and spinal cord, which generate eye movements, twitches of extremities, and presynaptic inhibition; BRF: bulbar reticular formation neurons that send tonic hyperpolarizing signals to motoneurons in spinal cord. As a consequence of these descending influences, sensory input and motor output are blocked at level of spinal cord. At level of forebrain, visual association and motor cortex neurons all receive time and phasic activation signals for nonspecific and specific thalamic relays.

References
 1. Aaronson, S. T., M. P. Biber, S. Rashed, and J. A. Hobson. Use of time lapse video tape recording in sleep research: method of study and sleep latency determination (Abstract). Sleep Res. 9: 118, 1980.
 2. Aghajanian, G. K., J. M. Cederbaum, and R. Y. Wang. Evidence for norepinephrine‐mediated collateral inhibition of the locus coeruleus neurons. Brain Res. 136: 570–577, 1977.
 3. Aghajanian, G. K., and H. J. Haigler. L‐Tryptophan as a selective histochemical marker for serotonergic neurons in single‐cell recording studies. Brain Res. 81: 364–372, 1974.
 4. Aghajanian, G. K., and R. Y. Wang. Habenular and other midbrain raphe afferents demonstrated by a modified retrograde tracing technique. Brain Res. 122: 229–242, 1977.
 5. Aghajanian, G. K., and R. Y. Wang. Physiology and pharmacology of central serotonergic neurons. In: Psychopharmacology: A Generation of Progress, edited by M. A. Lipton, A. DiMascio, and K. F. Killam. New York: Raven, 1978, p. 171–184.
 6. Aghajanian, G. K., R. Y. Wang, and J. Baraban. Serotonergic and nonserotonergic neurons of the dorsal raphe: reciprocal changes in firing induced by peripheral nerve stimulation. Brain Res. 153: 169–175, 1978.
 7. Agid, Y., F. Javoy, J. Glowinski, D. Bouvet, and C. Sotelo. Injection of 6‐hydroxydopamine into the substantia nigra of the rat. II. Diffusion and specificity. Brain Res. 58: 291–301, 1973.
 8. Ahlsén, G., and S. Lindström. Excitation of perigeniculate neurones via axon collaterals of principal cells. Brain Res. 236: 477–481, 1982.
 9. Ahlsén, G., S. Lindström, and F. S. Loo. Functional distinction of perigeniculate and thalamic reticular neurons of the cat. Exp. Brain Res. 46: 118–126, 1982.
 10. Ahlsén, G., S. Lindström, and E. Sybirska. Subcortical axon collaterals of principal cells in the lateral geniculate body of the cat. Brain Res. 156: 106–109, 1978.
 11. Akindle, M. O., J. I. Evans, and I. Oswald. Monoamine oxidase inhibitors, sleep and mood. Electroencephalogr. Clin. Neurophysiol. 29: 47–56, 1970.
 12. Albe‐Fessard, D., A. Levante, and Y. Lamour. Origin of spinothalamic and spinoreticular pathways in cats and monkeys. In: Advances in Neurology. International Symposium on Pain, edited by J. J. Bonica. New York: Raven, 1974, p. 157–166.
 13. Alger, B. E., and R. A. Nicoll. Epileptiform burst after‐hyperpolarization: calcium‐dependent potassium potential in hippocampal CA1 pyramidal cells. Science 210: 1122–1124, 1980.
 14. Allison, T. Cortical and subcortical evoked responses to central stimuli during wakefulness and sleep. Electroencephalogr. Clin. Neurophysiol. 18: 131–139, 1965.
 15. Allison, T., and G. D. Goff. Potentials evoked in somatosensory cortex to thalamo‐cortical radiation stimulation during waking, sleep and arousal from sleep. Arch. Ital. Biol. 106: 41–60, 1968.
 16. Altman, J., and S. A. Bayer. Development of the diencephalon in the rat. IV. Quantitative study of the time of origin of neurons and the internuclear chronological gradients in the thalamus. J. Comp. Neurol. 188: 455–472, 1979.
 17. Amaral, D. G., and H. M. Sinnamon. The locus coeruleus: neurobiology of a central noradrenergic nucleus. Prog. Neurobiol. 9: 147–196, 1977.
 18. Amassian, V. E., and R. V. De Vito. Unit activity in reticular formation and nearby structures. J. Neurophysiol. 17: 575–603, 1954.
 19. Amassian, V. E., and H. J. Waller. Spatio‐temporal patterns of activity in individual reticular neurons. In: Reticular Formation of the Brain, edited by H. H. Jasper, L. D. Proctor, R. S. Knighton, W. C. Noshay, and R. T. Costello. Boston, MA: Little, Brown, 1958, p. 69–108.
 20. Amatruda, T. T., III, D. A. Black, T. M. McKenna, R. W. McCarley, and J. A. Hobson. Sleep cycle control and cholinergic mechanisms: differential effects of carbachol at pontine brain stem sites. Brain Res. 98: 501–515, 1975.
 21. Andén, N. E., A. Carlsson, A. Dahlström, K. Fuxe, N. A. Hillarp, and K. Larsson. Demonstration and mapping out of nigro‐neostriatal dopamine neurons. Life Sci. 3: 523–531, 1964.
 22. Andén, N. E., A. Dahlstrom, K. Fuxe, K. Larsson, L. Olson, and V. Ungerstedt. Ascending monoamine neurons to the telencephalon and diencephalon. Acta Physiol. Scand. 67: 313–326, 1966.
 23. Andén, N., K. Fuxe, B. Hamberger, and T. Hökfelt. A quantitative study of the nigro‐neostriatal dopamine neuron system in rat. Acta Physiol. Scand. 67: 306–312, 1966.
 24. Andén, N. E., M. G. M. Jukes, A. Lundberg, and L. Vyklicky. The effect of DOPA on the spinal cord. Acta Physiol. Scand. 68: 322–336, 1966.
 25. Andersen, P., and S. A. Andersson. Physiological Basis of the Alpha Rhythm. New York: Appleton‐Century‐Crofts, 1968.
 26. Andersen, P., S. A. Andersson, and T. Lomo. Some factors involved in the thalamic control of spontaneous barbiturate spindles. J. Physiol. London 192: 257–281, 1967.
 27. Andersen, P., R. Dingledine, L. Gjerstad, I. A. Langmoen, and A. Laursen‐Mosfeldt. Two different responses of hippocampal pyramidal cells to application of gamma‐aminobutyric acid. J. Physiol. London 305: 279–296, 1980.
 28. Andersen, P., and J. C. Eccles. Inhibitory phasing of neuronal discharge. Nature London 196: 645–647, 1962.
 29. Andersen, P., J. C. Eccles, and T. A. Sears. The ventrobasal complex of the thalamus: types of cells, their responses and their functional organization. J. Physiol. London 174: 370–399, 1964.
 30. Andersen, P., and T. A. Sears. The role of inhibition in the phasing of spontaneous thalamocortical discharge. J. Physiol. London 173: 459–480, 1964.
 31. Anderson, M., and M. V. Salmon. Symptomatic cataplexy. J. Neurol. Neurosurg. Psychiatry 40: 186–191, 1977.
 32. Anderson, M. E., and M. Yoshida. Axonal branching patterns and location of nigrothalamic and nigrocollicular neurons in the cat. J. Neurophysiol. 43: 883–895, 1980.
 33. Andersson, S. A., and J. R. Manson. Rhythmic activity in the thalamus of the unanesthetized decorticate cat. Electroencephalogr. Clin. Neurophysiol. 31: 21–34, 1971.
 34. Angaut, P., and D. Bowsher. Ascending projections of the medial cerebellar (fastigial) nucleus: an experimental study in the cat. Brain Res. 24: 49–68, 1970.
 35. Angeleri, F., G. F. Marchesi, and A. Quattrini. Effects of chronic thalamic lesions on the electrical activity of the neocortex and on sleep. Arch. Ital. Biol. 107: 633–667, 1969.
 36. Apostol, G., and O. D. Creutzfeldt. Crosscorrelation between the activity of septal units and hippocampal EEG during arousal. Brain Res. 67: 65–75, 1974.
 37. Araki, T., and K. Endo. Short latency EPSPs of pyramidal tract cells evoked by stimulation of the centrum medianum‐parafascicular complex and the nucleus ventralis anterior of the thalamus. Brain Res. 113: 405–410, 1976.
 38. Arduini, A., G. Berlucchi, and P. Strata. Pyramidal activity during sleep and wakefulness. Arch. Ital. Biol. 101: 530–544, 1963.
 39. Asanuma, H., J. Fernandez, M. E. Scheibel, and A. B. Scheibel. Characteristics of projections from the nucleus ventralis lateralis to the motor cortex in the cats: an anatomical and physiological study. Exp. Brain Res. 20: 315–330, 1974.
 40. Aschoff, J. Circadian Clocks. Amsterdam: North‐Holland, 1965.
 41. Aschoff, J. Circadian rhythms in man. Science 148: 1427–1432, 1965.
 42. Aserinsky, E., and N. Kleitman. Regularly occurring periods of eye motility and concurrent phenomena during sleep. Science 118: 273–274, 1953.
 43. Aserinsky, E., and N. Kleitman. Two types of ocular motility occurring in sleep. J. Appl. Physiol. 8: 1–10, 1955.
 44. Ashby, W. R. Design for a Brain: The Origin of Adaptive Behaviour (2nd ed.). New York: Wiley, 1960.
 45. Ashby, W. R. Introduction to Cybernetics. New York: Science Eds., 1963.
 46. Aston‐Jones, G., and F. E. Bloom. Activity of norepinephrine‐containing locus coeruleus neurons in behaving rats anticipates fluctuations in the sleep‐waking cycle. J. Neurosci. 1: 876–886, 1981.
 47. Aston‐Jones, G., and F. E. Bloom. Norepinephrine‐containing locus coeruleus neurons in behaving rats exhibit pronounced responses to non‐noxious environmental stimuli. J. Neurosci. 1: 887–900, 1981.
 48. Astruc, J. Corticofugal fiber degeneration following lesions of area 8 (frontal eye field) in Macaca mulatta (Abstract). Anat. Rec. 148: 256, 1964.
 49. Axelrod, J. The pineal gland: a neurochemical transducer. Science 184: 1341–1348, 1974.
 50. Baghdoyan, H. A., A. P. Monaco, M. L. Rodrigo‐Angulo, F. Assens, R. W. McCarley, and J. A. Hobson. Microinjection of neostigmine into the pontine reticular formation of cats enhances desynchronized sleep signs. J. Pharmacol. Exp. Ther. 231: 173–180, 1984.
 51. Baghdoyan, H. A., M. L. Rodrigo‐Angulo, R. W. McCarley, and J. A. Hobson. Site‐specific enhancement and suppression of desynchronized sleep signs following cholinergic stimulation of three brainstem regions. Brain Res. 306: 39–52, 1984.
 52. Baker, T. L., and D. J. McGinty. Sleep apnea in hypoxic and normal kittens. Dev. Psychobiol. 12: 577–594, 1979.
 53. Baraitser, M., and J. K. Parkes. Genetic study of narcoleptic syndrome. J. Med. Genet. 15: 254–259, 1978.
 54. Barchas, J. D., W. C. Dement, D. Faull, A. S. Foutz, and R. B. Holzman. The concentration of amine metabolites in cerebrospinal fluid from normal and narcoleptic dogs. J. Physiol. London 296: 94–95, 1979.
 55. Barrionuevo, G., O. Benoit, and P. Tempier. Evidence for two types of firing pattern during the sleep‐waking cycle in the reticular thalamic nucleus of the cat. Exp. Neurol. 72: 486–501, 1981.
 56. Bartlett, J. R., and R. W. Doty, Sr. Influence of mesencephalic stimulation on unit activity in striate cortex of squirrel monkeys. J. Neurophysiol. 37: 642–652, 1974.
 57. Bartlett, J. R., R. W. Doty, J. Pecci‐Saavedra, and P. D. Wilson. Mesencephalic control of lateral geniculate nucleus in primates. III. Modifications with state of alertness. Exp. Brain Res. 18: 214–224, 1973.
 58. Bartolini, A., and G. Pepeu. Investigations into the acetylcholine output from the cerebral cortex of the cat in the presence of hyoscine. Br. J. Pharmacol. Chemother. 31: 66–74, 1967.
 59. Batsel, H. L. Spontaneous desynchronization in the chronic cat “cerveau isolé.” Arch. Ital. Biol. 102: 547–566, 1964.
 60. Baughman, R. W., and C. D. Gilbert. Aspartate and glutamate as possible neurotransmitters in the visual cortex. J. Neurosci. 1: 427–439, 1981.
 61. Baxter, B. L. Induction of both emotional behavior and a novel form of REM sleep by chemical stimulation applied to cat mesencephalon. Exp. Neurol. 23: 220–230, 1969.
 62. Belardetti, F., R. Borgia, and M. Mancia. Prosencephalic mechanisms of ECoG desynchronization in cerveau isolé cats. Electroencephahgr. Clin. Neurophysiol. 42: 213–225, 1977.
 63. Belin, M. F., M. Aguera, M. Tappaz, A. McRae‐Dequeurce, P. Bobillier, and J. F. Pujol. GABA‐accumulating neurons in the nucleus raphe dorsalis and periaqueductal gray in the rat: a biochemical and radioautographic study. Brain Res. 170: 279–297, 1979.
 64. Bell, C., G. Sierra, N. Buendia, and J. P. Segundo. Sensory properties of neurons in the mesencephalic reticular formation. J. Neurophysiol. 27: 961–987, 1964.
 65. Ben‐Ari, Y., R. Dingledine, I. Kanazawa, and J. S. Kelly. Inhibitory effects of acetylcholine on neurons in the feline nucleus reticularis thalami. J. Physiol. London 261: 647–671, 1976.
 66. Ben‐Ari, Y., K. Krnjević, W. Reinhardt, and N. Ropert. Intracellular observations on the disinhibitory action of acetylcholine in the hippocampus. Neuroscience 6: 2475–2484, 1981.
 67. Bénita, M., and H. Condé Intranuclear organization of the centre median nucleus of thalamus. J. Physiol. Paris 64: 561–582, 1972.
 68. Bentivoglio, M., G. Macchi, and A. Albanese. The cortical projections of the thalamic intralaminar nuclei, as studied in cat and rat with the multiple fluorescent retrograde tracing technique. Neurosci. Lett. 26: 5–10, 1981.
 69. Bentivoglio, M., G. Macchi, P. Rossini, and E. Tempesta. Brain stem neurons projecting to neocortex: an HRP study in the cat. Exp. Brain Res. 31: 489–498, 1978.
 70. Berger, H. Über das Elektrencephalogram des Menschen. Zweite Mitteilung. J. Psychol. Neurol. 40: 160–179, 1930.
 71. Berger, R. J. Oculomotor control: a possible function of REM sleep. Psychol. Rev. 76: 144–164, 1969.
 72. Berk, M. L., and J. A. Finkelstein. An autoradiographic determination of the efferent projections of the suprachiasmatic nucleus of the hypothalamus. Brain Res. 226: 1–13, 1981.
 73. Berkley, K. Spatial relationship between the termination of somatic sensory and motor pathways in the rostral brainstem of cats and monkeys. I. Ascending somatic sensory inputs to lateral diencephalon. J. Comp. Neurol. 193: 283–317, 1980.
 74. Berman, A. L. The Brain Stem of the Cat. Madison: Univ. of Wisconsin Press, 1968.
 75. Berman, A. L., and E. G. Jones. The Thalamus and Basal Telencephalon of the Cat. Madison: Univ. of Wisconsin Press, 1982.
 76. Berman, N. Connections of the pretectum in the cat. J. Comp. Neurol. 174: 227–254, 1977.
 77. Berrevoets, C. E., and H. G. J. M. Kuypers. Pericruciate cortical neurons projecting to brain stem reticular formation, dorsal column nuclei and spinal cord in the cat. Neurosci. Lett. 1: 257–262, 1975.
 78. Berridge, M. J. Modulation of nervous activity by cyclic nucleotides and calcium. In: The Neurosciences: Fourth Study Program, edited by F. O. Schmitt and F. G. Worden. Cambridge, MA: MIT Press, 1979, p. 873–889.
 79. Besset, A., A. Bonardet, M. Billiard, B. Descomps, A. C. De Paulet, and P. Passouant. Circadian patterns of growth hormone and Cortisol secretions in narcoleptic patients. Chronobiologia 6: 19–31, 1979.
 80. Besson, J. M., G. Guilbaud, M. Abdelmoumène, and A. Chaouch. Physiologie de la nociception. J. Physiol. Paris 78: 7–107, 1982.
 81. Bianchi, C., G. Spidelieri, P. Guandalini, S. Tangenelli, and L. Beani. Inhibition of acetylcholine outflow from the guinea‐pig cerebral cortex following locus coeruleus stimulation. Neurosci. Lett. 14: 97–100, 1979.
 82. Bishop, P. O. Properties of afferent synapses and sensory neurons in the lateral geniculate neurons. Int. Rev. Neurobiol. 6: 191–255, 1964.
 83. Bishop, P. O., W. Burke, and R. Davis. Single‐unit recording from antidromically activated optic radiation neurones. J. Physiol. London 162: 432–450, 1962.
 84. Bizzi, E., and D. C. Brooks. Functional connections between pontine reticular formation and lateral geniculate nucleus during deep sleep. Arch. Ital. Biol. 101: 666–680, 1963.
 85. Bland, B. H., P. Andersen, T. Ganes, and O. Sveen. Automated analysis of rhythmicity of physiologically identified hippocampal formation neurons. Exp. Brain Res. 38: 205–219, 1980.
 86. Bloch, V. Cerebral activation and memory fixation. Arch. Ital. Biol. 111: 577–590, 1973.
 87. Bloch, V., E. Hennevin, and P. Leconte. Relationship between paradoxical sleep and memory processes. In: Brain Mechanisms in Memory and Learning: From the Single Neuron to Man, edited by M. A. B. Frazier. New York: Raven, 1979, p. 329–343.
 88. Block, A. J., P. G. Boysen, J. W. Wynne, and L. A. Hunt. Sleep apnea, hypopnea and oxygen desaturation in normal subjects. A strong male predominance. N. Engl. J. Med. 300: 513–517, 1979.
 89. Bloom, F. E. The role of cyclic nucleotides in central synaptic function. Rev. Physiol. Biochem. Pharmacol. 74: 1–103, 1975.
 90. Bloom, F. E. Central noradrenergic systems: physiology and pharmacology. In: Psychopharmacology: A Generation of Progress, edited by M. A. Lipton, A. DiMascio, and K. F. Killam. New York: Raven, 1978, p. 131–141.
 91. Bloom, F. E. Chemical integrative processes in the central nervous system. In: The Neurosciences: Fourth Study Program, edited by F. O. Schmitt and F. G. Worden. Cambridge, MA: MIT Press, 1979, p. 51–58.
 92. Bloom, F. E. Chemical coding: modulation and level setting. Chairman's overview of Part IV. In: The Reticular Formation Revisited, edited by J. A. Hobson and M. A. B. Brazier. New York: Raven, 1980, p. 277–284.
 93. Bloom, F. E. How neurotransmitters may differentiate a cell's role in sensorimotor integration and behavioral state control. Neurosci. Res. Program Bull. 18: 19–21, 1980.
 94. Bloom, F. E., B. J. Hoffer, and G. R. Siggins. Studies on norepinephrine‐containing afferents to Purkinje cells of rat cerebellum. I. Localization to the fibers and their synapses. Brain Res. 25: 501–552, 1971.
 95. Bloom, F. E., B. J. Hoffer, and G. R. Siggins. Norepinephrine mediated cerebellar synapses: a model system for neuropsychopharmacology. Biol. Psychiatry 4: 157–177, 1972.
 96. Bobillier, P., S. Sequin, F. Petitjean, D. Salvert, M. Touret, and M. Jouvet. The raphe nuclei of cat brain stem: a topographical atlas of their efferent projections as revealed by autoradiography. Brain Res. 113: 449–486, 1976.
 97. Boivie, J. The termination of the spinothalamic tract in the cat. An experimental study with silver impregnation methods. Exp. Brain Res. 12: 331–353, 1971.
 98. Bolam, J. P., P. Somogyi, S. Totterdell, and A. D. Smith. A second type of striatonigral neuron: a comparison between retrogradely labelled and Golgi‐stained neurons at the light and electron microscopic level. Neuroscience 6: 2141–2157, 1981.
 99. Bouyer, J. J., M. F. Montaron, and A. Rougeul. Fast fronto‐parietal rhythms during combined focused attentive behaviour and immobility in cat: cortical and thalamic localizations. Electroencephahgr. Clin. Neurophysiol. 51: 244–252, 1981.
 100. Bowsher, D. Some afferent and efferent connections of the parafascicular‐center median complex. In: The Thalamas, edited by D. P. Purpura and M. D. Yahr. New York: Columbia Univ. Press, 1966, p. 99–108.
 101. Bowsher, D., and J. Westman. The gigantocellular reticular region and its spinal afferents: a light and electron microscope study in the cat. J. Anat. 106: 23–36, 1970.
 102. Boyle, R., and O. Pompeiano. Relation between cell size and response characteristics of vestibulospinal neurons to labyrinth and neck inputs. J. Neurosci. 1: 1052–1066, 1981.
 103. Bremer, F. Cerveau isolé et physiologie du commeil. C. R. Soc. Biol. 118: 1235–1241, 1935.
 104. Bremer, F. L'activité cérébrale au cours du sommeil et de la narcose. Contribution à l'étude du mécanisme du sommeil. Bull. Acad. R. Med. Belg. 4: 68–86, 1937.
 105. Bremer, F. Cerebral hypnogenic centers. Ann. Neurol. 2: 1–6, 1977.
 106. Bremer, F., and J. Chattonnet. Acétylcholine et cortex cérébral. Arch. Int. Physiol. 57: 106–109, 1949.
 107. Bremer, F., and N. Stoupel. Facilitation et inhibition des potentiels évoqués corticaux dans l'éveil cérébral. Arch. Int. Pharmacodyn. Ther. 67: 240–275, 1959.
 108. Bremer, F., and C. Terzuolo. 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. Arch. Int. Pharmacodyn. Ther. 62: 157–178, 1954.
 109. Brodal, A. The Reticular Formation of the Brain Stem: Anatomical Aspects and Functional Correlations. Edinburgh: Oliver & Boyd, 1957.
 110. Brodal, A., and G. F. Rossi. Ascending fibers in brain stem reticular formation of cat. Arch. Neurol. Psychiatry 74: 68–87, 1955.
 111. Bromberg, M. B., J. B. Penney, Jr., B. S. Stephenson, and A. B. Young. Evidence for glutamate as neurotransmitter of corticothalamic and corticorubral pathways. Brain Res. 215: 369–374, 1981.
 112. Brookhart, J. M. A study of cortico‐spinal activation of motoneurons. Res. Publ. Assoc. Res. Nerv. Ment. Dis. 30: 157–173, 1950.
 113. Brooks, D. C., and E. Bizzi. Brain‐stem electrical activity during deep sleep. Arch. Ital. Biol. 101: 648–665, 1963.
 114. Brooks, D. C., and M. D. Gershon. Eye movement potentials in the oculomotor and visual systems of the cat: a comparison of reserpine induced waves with those present during wakefulness and rapid eye movement sleep. Brain Res. 27: 223–239, 1971.
 115. Broughton, R. J. Sleep disorders: disorders of arousal? Science 159: 1070–1078, 1968.
 116. Bunney, B. S., and G. K. Aghajanian. The precise localization of nigral afferents in the rat as determined by a retrograde tracing technique. Brain Res. 117: 423–435, 1976.
 117. Burke, R. E., R. P. Dum, J. W. Fleshman, L. L. Glenn, A. Lev‐Tov, M. J. O'Donovan, and M. J. Pinter. An HRP study of the relation between cell size and motor unit type in cat ankle extensor motoneurons. J. Comp. Neurol. 209: 17–28, 1982.
 118. Burke, W., and A. M. Cole. Extraretinal influences on the lateral geniculate nucleus. Rev. Physiol. Biochem. Pharmacol. 80: 105–166, 1978.
 119. Burke, W., and A. J. Sefton. Inhibitory mechanisms in lateral geniculate nucleus of rat. J. Physiol. London 187: 231–246, 1966.
 120. Burns, B. D., and A. C. Webb. Mechanisms governing the relation between activity in the cerebral cortex and level of arousal. Proc. R. Soc. London Ser. B 214: 325–334, 1982.
 121. Buser, P., and F. E. Horvath. Thalamo‐caudate‐cortical relationships in synchronized activity. II. Further differentiation between spindle systems by cooling and lesions in the mesencephalon. Brain Res. 39: 43–60, 1972.
 122. Butcher, L. L., and J. Engel. Behavioral and biochemical effects on L‐dopa after peripheral decarboxylase inhibition. Brain Res. 15: 233–242, 1969.
 123. Büttner‐Ennever, J. A. Pathways from the pontine reticular formation to structures controlling horizontal and vertical eye movements in the monkey. In: Control of Gaze by Brainstem Neurons, edited by R. Baker and A. Berthoz. Amsterdam: Elsevier, 1977, p. 89–98.
 124. Buttner‐Ennever, J. A., and V. Henn. An autoradiographic study of the pathways from the pontine reticular formation involved in horizontal eye movements. Brain Res. 108: 155–164, 1976.
 125. Candia, O., E. Favale, A. Guissani, and G. F. Rossi. Blood pressure during natural sleep and during sleep induced by electrical stimulation of the brain stem reticular formation. Arch. Ital. Biol. 100: 216–233, 1962.
 126. Card, J. P., N. Brecha, H. J. Karten, and R. Y. Moore. Immunocytochemical localization of vasoactive intestinal polypeptide‐containing cells and processes in the suprachiasmatic nucleus of the rat: light and electron microscopic analysis. J. Neurosci. 1: 1289–1303, 1981.
 127. Carman, J. B., W. M. Cowan, and T. P. S. Powell. Cortical connexions of the thalamic reticular nucleus. J. Anat. 98: 587–598, 1964.
 128. Carter, D. A., and H. C. Fibiger. Ascending projections of presumed dopamine‐containing neurons in the ventral tegmentum of the rat as demonstrated by horseradish peroxidase. Neuroscience 2: 569–576, 1977.
 129. Carter, D. A., and H. C. Fibiger. The projections of the entopeduncular nucleus and globus pallidus in rat as demonstrated by autoradiography and horseradish peroxidase histochemistry. J. Comp. Neurol. 177: 113–124, 1978.
 130. Castaigne, P., A. Buge, R. Escourolle, and M. Masson. Ramollissement pédonculaire médian, tegmento‐thalamique avec ophthalmoplégie et hypersomnie. Rev. Neurol. 106: 357–367, 1962.
 131. Castiglione, A. J., M. C. Gallaway, and J. D. Coulter. Spinal projections from the midbrain in monkey. J. Comp. Neurol. 178: 329–346, 1978.
 132. Catsman‐Berrevoets, C. E., and H. G. J. M. Kuypers. A search for corticospinal collaterals to the thalamus and mesencephalon by means of multiple retrograde fluorescent tracers in cat and rat. Brain Res. 218: 15–33, 1981.
 133. Caviness, V. S., Jr., and D. O. Frost. Tangential organization of thalamic projections to the neocortex in the mouse. J. Comp. Neurol. 194: 335–367, 1980.
 134. Cederbaum, J. M., and G. K. Aghajanian. Afferent projections to the rat locus coeruleus as determined by a retrograde tracing technique. J. Comp. Neurol. 178: 1–16, 1978.
 135. Celesia, G. G., and H. H. Jasper. Acetylcholine released from cerebral cortex in relation to state of activation. Neurology 16: 1053–1064, 1966.
 136. Cesaro, P., J. Nguyen‐Legros, B. Berger, C. Alvarez, and D. Albe‐Fessard. Double labelling of branched neurons in the central nervous system of the rat by retrograde axonal transport of horseradish peroxidase and iron dextran complex. Neurosci. Lett. 15: 1–7, 1979.
 137. Cespuglio, R., H. Faradji, M. E. Gomez, and M. Jouvet. Single unit recordings in the nuclei raphe dorsalis and magnus during the sleep‐waking cycle of semi‐chronic prepared cats. Neurosci. Lett. 24: 133–138, 1981.
 138. Cespuglio, R., M. E. Gomez, E. Walker, and M. Jouvet. Effets du refroidissement et de la stimulation des noyaux du système du raphé sur les états de vigilance chez le chat. Electroencephalogr. Clin. Neurophysiol. 47: 289–308, 1979.
 139. Cespuglio, R., F. Rion, M. Buda, H. Faradji, F. Gonon, M. Jouvet, and J. F. Pujol. Mesure in vivo, par voltamétrie impulsionelle différentielle, du 5‐HIAA dans le striatum du rat. C. R. Acad. Sci. 290: 901–906, 1980.
 140. Chalmers, J. P., R. J. Baldessarini, and R. J. Wurtman. Effects of L‐Dopa on norepinephrine metabolism in the brain. Proc. Natl. Acad. Sci. USA 68: 662–666, 1971.
 141. Chase, M. H., S. H. Chandler, and Y. Nakamura. Intracellular determination of membrane potential of trigeminal motoneurons during sleep and wakefulness. J. Neurophysiol. 44: 349–358, 1980.
 142. Cheney, D. L., H. F. Le Feure, and G. Racagni. Choline acetyltransferase activity and mass fragmentographic measurement of acetylcholine in specific nuclei and tracts of rat brain. Neuropharmacology 14: 801–809, 1975.
 143. Chi, C. C., and J. P. Flynn. Neuroanatomic projections related to biting attack elicited from hypothalamus in cats. Brain Res. 35: 49–66, 1971.
 144. Chu, N. S., and F. E. Bloom. Norepinephrine‐containing neurons: changes in spontaneous discharge patterns during sleeping and waking. Science 179: 908–910, 1973.
 145. Chu, N. S., and F. E. Bloom. Activity patterns of catecholamine‐containing pontine neurons in the dorsolateral tegmentum of unrestrained cats. J. Neurobiol. 5: 527–544, 1974.
 146. Chu, N. S., and F. E. Bloom. The catecholamine‐containing neurons in the cat dorsolateral pontine tegmentum: distribution of the cell bodies and some axonal projections. Brain Res. 66: 1–21, 1974.
 147. Churukanov, V., B. Pollin, and D. Albe‐Fessard. Inhibition d'activités thalamiques par la stimulation du noyau inférieur du raphe chez le chat éveillé chronique. C. R. Acad. Sci. 283: 1651–1654, 1976.
 148. Clamann, H. P., and E. Hennemann. Electrical measurement of axon diameter and its use in relating motoneuron size to critical firing level. J. Neurophysiol. 39: 844–851, 1976.
 149. Clark, R. W., H. Boudoulas, S. F. Schaal, and H. S. Schmidt. Adrenergic hyperactivity and cardiac abnormality in primary disorders of sleep. Neurology 30: 113–119, 1980.
 150. Coenen, A. M. L., and A. J. H. Vendrik. Determination of the transfer ratio of cat's geniculate neurons through quasi‐intracellular recordings and the relation with the level of alertness. Exp. Brain Res. 14: 227–242, 1972.
 151. Cohen, B., and V. Henn. Unit activity in the pontine reticular formation associated with eye movement. Brain Res. 46: 403–410, 1972.
 152. Collier, B., and J. F. Mitchell. The central release of acetylcholine during consciousness and after brain lesions. J. Physiol. London 188: 83–99, 1967.
 153. Colonnier, M. L. The structural design of the neocortex. In: Brain and Conscious Experience, edited by J. C. Eccles. New York: Springer‐Verlag, 1966, p. 1–23.
 154. Colonnier, M., M. Steriade, and P. Landry. Selective resistance of sensory cells of the mesencephalic trigeminal nucleus to kainic acid‐induced lesions. Brain Res. 172: 552–556, 1979.
 155. Condé, F., and H. Condé Observations on the orthograde and retrograde transport of horseradish peroxidase in the cat. J. Hirnforsch. 20: 35–46, 1979.
 156. Coombs, J. S., J. C. Eccles, and P. Fatt. The electrical properties of the motoneurone membrane. J. Physiol. London 130: 291–325, 1955.
 157. Cordeau, J. P., J. De Champlain, and B. Jacks. Excitation and prolonged waking produced by catecholamines injected into the ventricular system of cats. Can. J. Physiol. Pharmacol. 49: 627–631, 1971.
 158. Coulter, J. D., B. K. Lester, and H. L. Williams. Reserpine and sleep. Psychopharmacologia 19: 134–147, 1971.
 159. Coyle, J. T., M. E. Molliver, and M. J. Kuhar. In situ injection of kainic acid: a new method for selectively lesioning neuronal cell bodies while sparing axons of passage. J. Comp. Neurol. 180: 301–324, 1978.
 160. Cramer, H., and W. Kuhlo. Effets des inhibiteurs de la mono‐aminoxydase sur le sommeil et l'électroencephalogramme chez l'homme. Acta Neurol. Belg. 67: 658–669, 1967.
 161. Crawford, J. M., and D. R. Curtis. Pharmacological studies on feline Betz cells. J. Physiol. London 186: 121–138, 1966.
 162. Creutzfeldt, O. D., S. Watanabe, and H. D. Lux. Relations between EEG phenomena and potentials of single cortical cells. I. Evoked responses after thalamic and epicortical stimulation. Electroencephalogr. Clin. Neurophysiol. 20: 1–18, 1966.
 163. Crow, T. J., J. F. Deakin, S. E. File, A. Longden, and S. Wendlandt. The locus coeruleus noradrenergic system—evidence against a role in attention, habituation, anxiety and motor activity. Brain Res. 155: 249–261, 1978.
 164. Crutcher, M. D., M. H. Branch, M. R. De Long, and A. P. Georgopoulos. Activity of zona incerta neurons in the behaving primate. Soc. Neurosci. Abstr. 6: 676, 1980.
 165. Cummings, J. P., and D. L. Felten. A raphe dendrite bundle in the rabbit medulla. J. Comp. Neurol. 183: 1–24, 1979.
 166. Dagnino, N., E. Favale, C. Loeb, and M. Manfredi. Sensory transmission in the geniculostriate system of the cat during natural sleep and arousal. J. Neurophysiol. 28: 443–456, 1965.
 167. Dagnino, N., E. Favale, M. Manfredi, A. Seitun, and A. Tartaglione. Tonic changes in excitability of thalamocortical neurons during the sleep‐waking cycle. Brain Res. 29: 354–357, 1971.
 168. Dahlström, A., and K. Fuxe. Evidence for the existence of monoamine neurons in the central nervous system. I. Demonstration of monoamines in the cell bodies of brain stem neurons. Acta Physiol. Scand. Suppl. 232: 1–55, 1964.
 169. Darke, P. G., and V. Jessen. Narcolepsy in a dog. Vet. Rec. 101: 117–118, 1977.
 170. Darwin, C. The Origin of Species by Means of Natural Selection. London: Murray, 1878.
 171. Dell, P. Some basic mechanisms of the translation of bodily needs into behaviour. In: Neurological Basis of Behaviour, edited by G. E. W. Wolstenholme and C. M. O'Connor. London: Churchill, 1957, p. 187–203.
 172. De Long, M. R. Activity of pallidal neurons during movement. J. Neurophysiol. 34: 414–427, 1971.
 173. De Long, M. R. Activity of basal ganglia neurons during movement. Brain Res. 40: 127–135, 1972.
 174. Delorme, F., M. Jeannerod, and M. Jouvet. Effets remarquables de la réserpine sur l'activité EEG phasique pontogeniculo‐occipitale. C. R. Soc. Biol. 159: 900–903, 1965.
 175. Delphs, J. R., and M. A. Dichter. Effects of somastatin on mammalian cortical neurons in culture: physiological actions and unusual dose response characteristics. J. Neurosci. 3: 1176–1188, 1983.
 176. Dement, W. The occurrence of low voltage, fast, electroencephalogram patterns during behavioral sleep in the cat. Electroencephalogr. Clin. Neurophysiol. 10: 291–296, 1958.
 177. Dement, W. C., S. Hendriksen, B. L. Jacobs, and M. M. Mitler. Biogenic amines, phasic events, and behavior. In: Pharmacology and the Future of Man. Brain, Nerves, and Synapses, edited by F. E. Bloom and G. H. Acheson. New York: Karger, 1973, vol. 4, p. 74–89.
 178. Dement, W., and N. Kleitman. Cyclic variations in EEG during sleep and their relation to eye movements, body mobility and dreaming. Electroencephalogr. Clin. Neurophysiol. 9: 673–690, 1957.
 179. Dement, W., and N. Kleitman. The relation of eye movements during sleep to dream activity: an objective method for the study of dreaming. J. Exp. Psychol. 53: 89–97, 1957.
 180. De Montigny, C., and J. P. Lund. Microiontophoretic study of the action of kainic acid and putative transmitters in the rat mesencephalic trigeminal nucleus. Neuroscience 5: 1621–1628, 1980.
 181. Deniau, J. M., A. M. Thierry, and J. Feger. Electrophysiological identication of mesencephalic ventromedial tegmental (VMT) neurons projecting to the frontal cortex, septum and nucleus accumbens. Brain Res. 189: 315–326, 1980.
 182. Descarries, L., A. Beaudet, K. C. Watkins, and S. Garcia. The serotonin neurons in nucleus raphe dorsalis of adult rat (Abstract). Anat. Rec. 193: 520, 1979.
 183. Descarries, L., K. C. Watkins, and Y. Lapierre. Noradrenergic axon terminals in the cerebral cortex of rat. III. Topometric ultrastructural analysis. Brain Res. 133: 197–222, 1977.
 184. Deschěnes, M., M. Paradis, J. P. Roy, and M. Steriade. Electrophysiology of neurons of lateral thalamic nuclei in cat: resting properties and burst discharges. J. Neurophysiol. 51: 1196–1219, 1984.
 185. Deschěnes, M., J. P. Roy, and M. Steriade. Thalamic bursting mechanism: an inward slow current revealed by membrane hyperpolarization. Brain Res. 239: 289–293, 1982.
 186. Deschěnes, M., J. P. Roy, and M. Steriade. Nature of 7–14 Hz rhythmic hyperpolarizations in thalamic relay neurons. Soc. Neurosci. Abstr. 9: 678, 1983.
 187. Desmedt, J. E., and J. Debecker. Wave form and neural mechanism of the decision P350 elicited without pre‐stimulus CNV or readiness potential in random sequences of near‐threshold auditory clicks and finger stimuli. Electroencephalogr. Clin. Neurophysiol. 47: 648–670, 1979.
 188. Desmedt, J. E., and D. Robertson. Differential enhancement of early and late components of the cerebral somatosensory evoked potentials during forced‐pace cognitive tasks in man. J. Physiol. London 271: 761–782, 1977.
 189. Dewan, E. M. The Programming (P) Hypothesis for REMs. Washington, DC: Office of Aerospace Res., USAF, 1969. (Phys. Sci. Res. Paper 388.)
 190. Dierickx, K., and F. Vandesande. Immunocytochemical localization of somatostatin‐containing neurons in the rat hypothalamus. Cell Tissue Res. 201: 349–359, 1979.
 191. Dingledine, R., and J. S. Kelly. Brain stem stimulation and the acetylcholine‐evoked inhibition of neurones in the feline nucleus reticularis thalami. J. Physiol. London 271: 135–154, 1977.
 192. Dismukes, R. K. New concepts of molecular communication among neurons. Behav. Brain Sci. 2: 409–448, 1979.
 193. Divac, I. Magnocellular nuclei of the basal forebrain project to neocortex, brain stem, and olfactory bulb. Review of some functional correlates. Brain Res. 93: 385–398, 1975.
 194. Domer, F. R., and W. Feldberg. Some central actions of adrenaline and noradrenaline when administered into the cerebral ventricles. In: Adrenergic Mechanisms, edited by G. E. W. Wolstenholme and M. J. O'Connor. London: Churchill, 1960, p. 386–392. (Ciba Found. Symp.)
 195. Domino, E. F., and M. Stawiski. Modification of the cat sleep cycle by hemicholinium‐3, a cholinergic antisynthesis agent. Res. Commun. Chem. Pathol. Pharmacol. 2: 461–467, 1971.
 196. Domino, E. F., and K. Yamamoto. Nicotine effect on the sleep cycle of the cat. Science 150: 637–638, 1965.
 197. Domino, E. F., K. Yamamoto, and A. T. Dren. Role of cholinergic mechanisms in state of wakefulness and sleep. In: Progress in Brain Research. Anticholinergic Drugs, edited by P. B. Bradley and M. Fink. Amsterdam: Elsevier, 1968, vol. 28, p. 113–133.
 198. Donoghue, J. P., and F. F. Ebner. The laminar distribution and ultrastructure of fibers projecting from three thalamic nuclei to the somatic sensory‐motor cortex of the opossum. J. Comp. Neurol. 198: 389–420, 1981.
 199. Doty, R. W. Electrical stimulation of the brain in behavioral context. Annu. Rev. Psychol. 20: 289–320, 1969.
 200. Dubin, M. W., and B. G. Cleland. Organization of visual inputs to interneurons of lateral geniculate nucleus of the cat. J. Neurophysiol. 40: 410–427, 1977.
 201. Dumont, S., and P. Dell. Facilitation réticulaire des mécanismes visuels corticaux. Electroencephalogr. Clin. Neurophysiol. 12: 769–796, 1958.
 202. Dunleavy, D. L. F., V. Brezinova, I. Oswald, A. W. MacLean, and M. Tinkler. Changes during weeks in effects of tricyclic drugs on the human sleeping brain. Br. J. Psychiatry. 120: 663–672, 1972.
 203. Eccles, J. C. The Physiology of Synapses. Berlin: Springer‐Verlag, 1964.
 204. Eccles, J. C., R. A. Nicoll, D. W. F. Schwarz, H. Táboríková, and T. J. Willey. Reticulospinal neurons with and without monosynaptic inputs from cerebellar nuclei. J. Neurophysiol. 38: 513–530, 1975.
 205. Eccles, J. C., R. A. Nicoll, H. Táboríková, and T. J. Willey. Medial reticular neurons projecting rostrally. J. Neurophysiol. 38: 531–538, 1975.
 206. Edley, S. M., and A. M. Graybiel. The afferent and efferent connections of the feline nucleus tegmenti pedunculopontinus, pars compacta. J. Comp. Neurol. 217: 187–215, 1983.
 207. Edwards, S. B. Autoradiographic studies of the projections of the midbrain reticular formation: descending projections of nucleus cuneiformis. J. Comp. Neurol. 161: 341–358, 1975.
 208. Edwards, S. B. The deep cell layers of the superior colliculus: their reticular characteristics and structural organization. In: The Reticular Formation Revisited, edited by J. A. Hobson and M. A. B. Brazier. New York: Raven, 1980, p. 193–209.
 209. Edwards, S. B., and J. S. De Olmos. Autoradiographic studies of the projections of the midbrain reticular formation: ascending projections of nucleus cuneiformis. J. Comp. Neurol. 165: 417–432, 1976.
 210. Egan, T. M., G. Henderson, R. A. North, and J. T. Williams. Noradrenaline‐mediated synaptic inhibition in rat locus coeruleus neurones. J. Physiol. London 345: 477–488, 1983.
 211. Emson, P.C., G. Paxinos, G. Le Gal La Salle, Y. Ben‐Ari, and A. Silver. Choline acetyltransferase and acetylcholinesterase containing projections from the basal forebrain to the amygdaloid complex of the rat. Brain Res. 165: 271–282, 1979.
 212. Endo, K., T. Araki, and K. Ito. Short latency EPSPs and incrementing PSPs of pyramidal tract cells evoked by stimulation of the nucleus centralis lateralis of the thalamus. Brain Res. 132: 541–546, 1977.
 213. Evarts, E. V. Effects of sleep and waking on spontaneous and evoked discharge of single units in visual cortex. Federation Proc. Suppl. 4: 828–837, 1960.
 214. Evarts, E. V. Temporal patterns of discharge of pyramidal tract neurons during sleep and waking in the monkey. J. Neurophysiol. 27: 152–171, 1964.
 215. Evarts, E. V. Relation of discharge frequency to conduction velocity in pyramidal tract neurons. J. Neurophysiol. 28: 216–228, 1965.
 216. Everett, G. M., and J. W. Borchering. L‐Dopa: effect on concentrations of dopamine, norepinephrine and serotonin in brains of mice. Science 168: 849–850, 1970.
 217. Facon, E., M. Steriade, and N. Wertheim. Hypersomnie prolongée engendrée par des lésions bilatérales du système activateur médial. Le syndrome thrombotique de la bifurcation du tronc basilaire. Rev. Neurol. 98: 117–133, 1958.
 218. Falck, B., N. A. Hillarp, G. Thieme, and A. Torp. Fluorescence of catecholamines and related compounds condensed with formaldehyde. J. Histochem. Cytochem. 10: 348–354, 1962.
 219. Fallon, J. H. Collateralization of monoamine neurons: mesotelencephalic dopamine projections to caudate, septum, and frontal cortex. J. Neurosci. 1: 1361–1368, 1981.
 220. Fallon, J. H., and R. Y. Moore. Catecholamine innervation of the basal forebrain. IV. Topography of the dopamine projection to the basal forebrain and neostriatum. J. Comp. Neurol. 180: 545–580, 1978.
 221. Farber, J., J. D. Miller, P. Gatz, D. C. German, and H. P. Roffwarg. Unit activity of CA and non‐CA neurons in midbrain (DA) and hindbrain (NE) in relation to sleep‐wake states (Abstract). Sleep Res. 10: 27, 1981.
 222. Favale, E., C. Loeb, M. Manfredi, and G. Sacco. Somatic afferent transmission and cortical responsiveness during natural sleep and arousal in the cat. Electroencephalogr. Clin. Neurophysiol. 18: 354–368, 1965.
 223. Fedina, L., A. Lundberg, and L. Vyklický The effect of a noradrenaline liberator (4, α‐dimethyl‐meta‐tyramine) on reflex transmission in spinal cats. Acta Physiol. Scand. 83: 495–504, 1971.
 224. Feldberg, W., and S. L. Sherwood. Injections of drugs into the lateral ventricle of the cat. J. Physiol. London 123: 148–167, 1954.
 225. Feldman, S. M., and H. J. Waller. Dissociation of electrocortical activation and behavioural arousal. Nature London 196: 1320–1322, 1962.
 226. Ferster, D., and S. Levay. The axonal arborizations of lateral geniculate neurons in the striate cortex of the cat. J. Comp. Neurol. 182: 923–944, 1978.
 227. Fertziger, A. P., and D. P. Purpura. Diphasic‐PSPs during maintained activity of cat lateral geniculate neurons. Brain Res. 33: 463–467, 1971.
 228. Filion, M., and C. Harnois. A comparison of projections of entopeduncular neurons to the thalamus, the midbrain and the habenula in the cat. J. Comp. Neurol. 181: 763–780, 1978.
 229. Filion, M., Y. Lamarre, and J. P. Cordeau. Neuronal discharges of the ventrolateral nucleus of the thalamus during sleep and wakefulness in the cat. II. Evoked activity. Exp. Brain Res. 12: 499–508, 1971.
 230. Fishbein, W., and B. M. Gutwein. Paradoxical sleep and memory storage processes. Behav. Biol. 19: 425–464, 1977.
 231. Flicker, C., R. W. McCarley, and J. A. Hobson. Aminergic neurons: state control and plasticity in three model systems. Cell. Mol. Neurobiol. 1: 123–166, 1981.
 232. Foote, S. L. Compensatory changes in REM sleep time of cats during ad libitum sleep and following brief REM sleep deprivation. Brain Res. 54: 261–276, 1973.
 233. Foote, S. L., G. Aston‐Jones, and F. E. Bloom. Impulse activity of locus coeruleus neurons in awake rats and monkeys is a function of sensory stimulation and arousal. Proc. Natl. Acad. Sci. USA 77: 3033–3037, 1980.
 234. Foote, S. L., and F. E. Bloom. Activity of norepinephrine‐containing locus coeruleus neurons in the unanesthetized squirrel monkey. In: Catecholamines: Basic and Clinical Frontiers, edited by E. Usdin, I. J. Kopin, and J. Barchas. New York: Pergamon, 1979, p. 625–627.
 235. Foote, S. L., F. E. Bloom, and G. Aston‐Jones. Nucleus locus ceruleus: new evidence of anatomical and physiological specificity. Physiol. Rev. 63: 844–914, 1983.
 236. Foote, S. L., R. Freedman, and A. P. Olivier. Effects of putative neurotransmitters on neuronal activity in monkey auditory cortex. Brain Res. 86: 229–242, 1975.
 237. Foster, J. A. Intracortical origin of recruiting responses in the cat cortex. Electroencephalogr. Clin. Neurophysiol. 48: 639–653, 1980.
 238. Foulkes, D. The Psychology of Sleep. New York: Scribner, 1966.
 239. Fourment, A., and J. C. Hirsch. Single‐unit discharges in the dorsolateral thalamus of behaving cats: spontaneous activity. Exp. Neurol. 65: 1–15, 1979.
 240. Fram, D. H., D. L. Murphy, F. K. Goodwin, H. Keith, H. Brodie, W. E. Bunney, Jr., and F. Snyder. L‐Dopa and sleep in depressed patients. Psychophysiology 7: 316–317, 1970.
 241. Frigyesi, T. L. Intracellular recordings from neurons in dorsolateral thalamic reticular nucleus during capsular, basal ganglia and midline thalamic stimulation. Brain Res. 48: 157–172, 1972.
 242. Frigyesi, T. L., and R. Schwartz. Cortical control of thalamic sensorimotor relay activities in the cat and the squirrel monkey. In: Corticothalamic Projections and Sensorimotor Activities, edited by T. Frigyesi, E. Rinvik, and M. D. Yahr. New York: Raven, 1972, p. 161–191.
 243. Frost, D. O., and V. S. Caviness Jr. Radial organization of thalamic projections to the neocortex in the mouse. J. Comp. Neurol. 194: 369–393, 1980.
 244. Fujita, Y., and T. Sato. Intracellular records from hippocampal pyramidal cells in rabbit during theta rhythm activity. J. Neurophysiol. 27: 1011–1025, 1964.
 245. Fukuda, Y., and K. Iwama. Inhibition des interneurones du corps genouillé latéral par activation de la formation réticulée. Brain Res. 18: 548–551, 1970.
 246. Fuller, C. A., R. Lydic, F. M. Sulzman, H. E. Albers, B. Tepper, and M. C. Moore‐Ede. Circadian rhythm of body temperature persists after suprachiasmatic lesions in the squirrel monkey. Am. J. Physiol. 241 (Regulatory Integrative Comp. Physiol. 10): R385–R391, 1981.
 247. Fuller, J. H. Brain stem reticular units: some properties of the course and origin of the ascending trajectory. Brain Res. 83: 349–367, 1975.
 248. Fuller, J. H., and J. D. Schlag. Determination of antidromic excitation by the collision test: problems of interpretation. Brain Res. 112: 283–298, 1976.
 249. Fuster, J. M., and A. A. Uyeda. Facilitation of tachistoscopic performance by stimulation of midbrain tegmental points in the monkey. Exp. Neurol. 6: 384–406, 1962.
 250. Fuxe, K., T. Hökfelt, and U. Ungestedt. Morphological and functional aspects of central monoamine neurons. Int. Rev. Neurobiol. 13: 93–126, 1970.
 251. Galambos, R., G. Sheatz, and V. G. Vernier. Electrophysiological correlates of a conditioned response in cats. Science 123: 376–377, 1956.
 252. Gallagher, D. W., and A. Pert. Efferents to the brain stem nuclei (brain stem raphe, nucleus reticularis pontis caudalis and nucleus gigantocellularis) in the rat as demonstrated by microiontophoretically applied horseradish peroxidase. Brain Res. 144: 257–275, 1978.
 253. Garey, L. J., and J. P. Hornung. The use of ibotenic acid lesions for light and electron microscopic study of anterograde degeneration in the visual pathway of the cat. Neurosci. Lett. 19: 117–123, 1980.
 254. George, R., W. L. Haslett, and D. J. Jenden. A cholinergic mechanism in the pontine reticular formation: induction of paradoxical sleep. Int. J. Neuropharmacol. 3: 541–552, 1964.
 255. Giesler, G. J., Jr., R. P. Yezierski, K. D. Gerhart, and W. D. Willis. Spinothalamic tract neurons that project to medial and/or lateral thalamic nuclei: evidence for a physiologically novel population of spinal cord neurons. J. Neurophysiol. 46: 1285–1308, 1981.
 256. Gillin, J. G., R. M. Post, R. J. Wyatt, F. K. Goodwin, F. Snyder, and W. Bunney Jr. REM inhibitory effect of L‐dopa infusion during human sleep. Electroencephalogr. Clin. Neurophysiol. 35: 181–186, 1973.
 257. Glenn, L. L., and W. C. Dement. Membrane potential, synaptic activity, and excitability of hindlimb motoneurons during wakefulness and sleep. J. Neurophysiol. 46: 839–854, 1981.
 258. Glenn, L. L., J. Hada, J. P. Roy, M. Deschěnes, and M. Steriade. Anterograde tracer and field potential analysis of the neocortical layer I projection from nucleus ventralis medialis of the thalamus in cat. Neuroscience 7: 1861–1877, 1982.
 259. Glenn, L. L., and M. Steriade. Discharge rate and excitability of cortically projecting neurons in the intralaminar thalamic nuclei during waking and sleep states. J. Neurosci. 2: 1387–1404, 1982.
 260. Goebel, H. H., A. Komatsuzaki, M. B. Bender, and B. Cohen. Lesions of the pontine tegmentum and conjugate gaze paralysis. Arch. Neurol. 24: 431–440, 1971.
 261. Goldberg, M. P., D. Riew, E. Vivaldi, R. W. McCarley, and J. A. Hobson. Activation of muscarinic receptors in the pontine tegmentum can trigger REM sleep: effects of bethanechol (Abstract). Sleep Res. 10: 81, 1981.
 262. Graham, J., and N. Berman. Origins of the pretectal and tectal projections to the central lateral nucleus in the cat. Neurosci. Lett. 26: 209–214, 1981.
 263. Grantyn, A., and R. Grantyn. Axonal patterns and sites of termination of cat superior colliculus neurons projecting in the tecto‐bulbo‐spinal tract. Exp. Brain Res. 46: 243–256, 1982.
 264. Grantyn, A., M. Mancia, G. Broggi, and M. Margnelli. Intralaminar thalamic influences on bulbo‐pontine and mesencephalic neurones as revealed by intracellular recording. Brain Res. 33: 223–226, 1971.
 265. Grantyn, R., M. Margnelli, M. Mancia, and A. Grantyn. Postsynaptic potentials in the mesencephalic and ponto‐medullary reticular regions underlying descending limbic influences. Brain Res. 56: 107–121, 1973.
 266. Graybiel, A. M. Direct and indirect preoculomotor pathways of the brainstem: an autoradiographic study of the pontine reticular formation in the cat. J. Comp. Neurol. 175: 37–78, 1977.
 267. Graybiel, A. M., and R. P. Elde. Somatostatin‐like immunoreactivity characterizes neurons of the nucleus reticularis thalami in the cat and monkey. J. Neurosci. 3: 1308–1321, 1983.
 268. Graybiel, A. M., and E. A. Hartwieg. Some afferent connections of the oculomotor complex in the cat: an experimental study with tracer techniques. Brain Res. 81: 543–551, 1974.
 269. Green, J. D., and A. Arduini. Hippocampal electrical activity in arousal. J. Neurophysiol. 17: 533–557, 1954.
 270. Grofova, I. The identification of striatal and pallidal neurons projecting to substantia nigra. An experimental study by means of retrograde axonal transport of horseradish peroxidase. Brain Res. 91: 286–291, 1975.
 271. Grofova, I., O. P. Ottersen, and E. Rinvik. Mesencephalic and diencephalic afferents to the superior colliculus and periaqueductal gray substance demonstrated by retrograde axonal transport of horseradish peroxidase in the cat. Brain Res. 146: 205–220, 1978.
 272. Grossman, S. P., D. Dacey, A. E. Halaris, T. Collier, and A. Routtenberg. Aphagia and adipsia after preferential destruction of nerve cell bodies in hypothalamus. Science 202: 537–539, 1978.
 273. Gücer, G. The effect of sleep upon the transmission of afferent activity in the somatic afferent system. Exp. Brain Res. 34: 287–298, 1979.
 274. Guilleminault, C., R. L. Ariagno, L. S. Forno, L. Nagel, R. Baldwin, and M. Owen. Obstructive sleep apnea and near miss for SIDS. 1. Report of an infant with sudden death. Pediatrics 63: 837–843, 1979.
 275. Guilleminault, C., R. Ariagno, R. Korobkin, L. Nagel, R. Baldwin, S. Coons, and M. Owen. Mixed and obstructive sleep apnea and near miss for sudden infant death syndrome. 2. Comparison of near miss and normal control infants by age. Pediatrics 64: 882–891, 1979.
 276. Guilleminault, C., and W. C. Dement. Sleep Apnea Syndromes. New York: Liss, 1978.
 277. Guilleminault, C., F. L. Eldridge, A. Tilkian, F. B. Simmons, and W. C. Dement. Sleep apnea syndrome due to upper airway obstruction: a review of 25 cases. Arch. Intern. Med. 137: 296–300, 1977.
 278. Hajdu, F., G. Somogyi, and T. Tömböl. Neuronal and synaptic arrangement in the lateralis posterior‐pulvinar complex of the thalamus in the cat. Brain Res. 73: 89–104, 1974.
 279. Hammer, R. P., Jr., R. D. Lindsay, and A. B. Scheibel. Development of the brain stem reticular core: an assessment of dendritic state and configuration in the perinatal rat. Dev. Brain Res. 7: 179–190, 1981.
 280. Harding, B. N., and T. P. S. Powell. An electron microscopic study of the centre‐median and ventrolateral nuclei of the thalamus in the monkey. Proc. R. Soc. London 279: 357–412, 1977.
 281. Hartmann, E., and T. J. Bridwell. Effects of AMPT, L‐dopa and L‐tryptophan on sleep in the rat (Abstract). Psychophysiology 7: 313, 1970.
 282. Hartmann, E., T. J. Bridwell, and J. J. Schildkraut. α‐Methylparatyrosine and sleep in the rat. Psychopharmacologia 21: 157–164, 1971.
 283. Hartmann, E., and G. Zwilling. Effect of α‐ and β‐blockers on sleep patterns in the rat (Abstract). Sleep Res. 1: 54, 1972.
 284. Hartmann, E., G. Zwilling, and S. List. Effects of an alpha‐adrenergic blocker on sleep in the rat (Abstract). Sleep Res. 2: 58, 1973.
 285. Hartmann, E., G. Zwilling, V. Pochay, and R. Chung. Intraventricular administration of neurotransmitters: effects on sleep (Abstract). Sleep Res. 3: 41, 1974.
 286. Hartse, K. M., S. F. Eisenhart, B. M. Bergmann, and A. Rechtschaffen. Ventral hippocampus spikes during sleep, wakefulness and arousal in the cat. Sleep 1: 231–246, 1979.
 287. Hashikawa, T., and K. Kawamura. Identification of cells of origin of tectopontine fibers in the cat superior colliculus: an experimental study with the horseradish peroxidase method. Brain Res. 130: 65–79, 1977.
 288. Hattori, T., H. C. Fibiger, and P. L. McGeer. Demonstration of a pallido‐nigral projection innervating dopaminergic neurons. J. Comp. Neurol. 162: 487–504, 1975.
 289. Hazlett, J. C., C. R. Dutta, and C. A. Fox. The neurons in the centromedian‐parafascicular complex of the monkey (Macaco mulatta): a Golgi study. J. Comp. Neurol. 168: 41–74, 1976.
 290. Hazra, J. Effect of hemicholinium‐3 on slow wave and paradoxical sleep of cat. Eur. J. Pharmacol. 11: 395–397, 1970.
 291. Hendry, S. H. C., and E. G. Jones. Sizes and distributions of intrinsic neurons incorporating tritiated GABA in monkey sensory‐motor cortex. J. Neurosci. 1: 390–408, 1981.
 292. Hendry, S. H. C., E. G. Jones, and J. Graham. Thalamic relay nuclei for cerebellar and certain related fiber systems in the cat. J. Comp. Neurol. 185: 679–714, 1979.
 293. Henley, K., and A. R. Morrison. A re‐evaluation of the effects of lesions of the pontine tegmentum and locus coeruleus on phenomena of paradoxical sleep in the cat. Acta Neurobiol. Exp. 34: 215–232, 1974.
 294. Henn, V., V. Büttner, and J. Büttner‐Ennever. Supranucleus organization of the oculomotor system—anatomical and physiological investigations. In: Disorders of Ocular Motility: Neurophysiological and Clinical Aspects, edited by G. Kommerell. Munich, West Germany: Bermann, 1978, p. 129–143.
 295. Henneman, E., G. Somjen, and D. O. Carpenter. Functional significance of cell size in spinal motoneurons. J. Neurophysiol. 28: 560–580, 1965.
 296. Henriksen, S. J., B. L. Jacobs, and W. C. Dement. Dependence of REM sleep PGO waves on cholinergic mechanisms. Brain Res. 48: 412–416, 1972.
 297. Henriksen, S., B. Jacobs, W. Dement, and J. Barchas. Catecholamine mechanisms: their presumptive role in the generation of REM sleep PGO waves. In: Frontiers in Catecholamine Research, edited by E. Usdin and S. H. Snyder. Oxford, UK: Pergamon, 1973, p. 759–762.
 298. Herkenham, M. The afferent and efferent connections of the ventromedial thalamic nucleus in the rat. J. Comp. Neurol. 183: 487–518, 1979.
 299. Herkenham, M. Laminar organization of thalamic projections to the rat neocortex. Science 207: 532–535, 1980.
 300. Herman, J. H., H. P. Roffwarg, C. J. Rosenmann, and E. J. Tauber. Binocular depth perception following REM deprivation or awake state visual deprivation. Psychophysiology 17: 236–242, 1980.
 301. Herman, Z. S. The effects of noradrenaline on rat's behavior. Psychopharmacologia 16: 369–374, 1970.
 302. Hernandez‐Peon, R. A cholinergic hypnogenic limbic forebrain‐hindbrain circuit. In: Neurophysiologie des états du sommeil, edited by M. Jouvet. Lyons, France: CNRS, 1965, p. 63–88.
 303. Hersch, S. M., and E. L. White. Thalamocortical synapses with corticothalamic projection neurons in mouse SmI cortex: electron microscopic demonstration of a monosynaptic feedback loop. Neurosci. Lett. 24: 207–210, 1981.
 304. Higuchi, T., Y. Takahashi, K. Takahashi, Y. Niimi, and A. Miyasita. Twenty‐four‐hour secretory patterns of growth hormones, prolactin, and Cortisol in narcolepsy. J. Clin. Endocrinol. Metab. 49: 197–204, 1979.
 305. Hirsch, J. C., A. Fourment, and M. E. Marc. Electrophysiological study of the peripeniculate region during natural sleep of the cat. Exp. Neurol. 2: 436–454, 1982.
 306. Hirsch, J. C., A. Fourment, and M. E. Marc. Sleep‐related variations of membrane potential in the lateral geniculate body relay neurons of the cat. Brain Res. 259: 308–312, 1983.
 307. Hishikawa, T., K. Nakai, H. Ida, and Z. Kaneko. The effect of imipramine, desmethylimipramine and chlorpromazine on the sleep‐wakefulness cycle of the cat. Electroencephalogr. Clin. Neurophysiol. 19: 518–521, 1965.
 308. Hobson, J. A. The effect of chronic brain stem lesions on cortical and muscular activity during sleep and waking in the cat. Electroencephalogr. Clin. Neurophysiol. 19: 41–62, 1965.
 309. Hobson, J. A. Sleep after exercise. Science 162: 1503–1505, 1968.
 310. Hobson, J. A. The cellular basis of sleep cycle control. In: Advances in Sleep Research, edited by E. D. Weitzman. New York: Spectrum, 1974, vol. 1, p. 217–250.
 311. Hobson, J. A. Dreaming sleep attacks and desynchronized sleep enhancement. Arch. Gen. Psychiatry 32: 1421–1424, 1975.
 312. Hobson, J. A. The sleep‐dream cycle: a neurobiological rhythm. In: Pathobiology Annual, edited by H. L. Ioachim. New York: Appleton‐Century‐Crofts, 1975, p. 369–703.
 313. Hobson, J. A. The Ethology of Sleep. Manati, Puerto Rico: Roche, 1979.
 314. Hobson, J. A., and M. A. B. Brazier (editors). The Reticular Formation Revisited. New York: Raven, 1979.
 315. Hobson, J. A., F. Goldfrank, and F. Snyder. Respiration and mental activity in sleep. J. Psychiatr. Res. 3: 79–90, 1965.
 316. Hobson, J. A., R. W. McCarley, R. Freedman, and R. T. Pivik. Time course of discharge rate changes by cat pontine brain stem neurons during sleep cycle. J. Neurophysiol. 37: 1297–1309, 1974.
 317. Hobson, J. A., R. W. McCarley, and J. P. Nelson. Location and spike‐train characteristics of cells in anterodorsal pons having selective decreases in discharge rates during desynchronized sleep. J. Neurophysiol. 50: 770–783, 1983.
 318. Hobson, J. A., R. W. McCarley, R. T. Pivik, and R. Freedman. Selective firing by cat pontine brain stem neurons in desynchronized sleep. J. Neurophysiol. 37: 497–511, 1974.
 319. Hobson, J. A., R. W. McCarley, and P. W. Wyzinski. Sleep cycle oscillation: reciprocal discharge by two brainstem neuronal groups. Science 189: 55–58, 1975.
 320. Hobson, J. A., and A. B. Scheibel. The brainstem core: sensorimotor integration and behavioral state control. Neurosci. Res. Program Bull. 18: 1–173, 1980.
 321. Hobson, J. A., T. Spagna, and R. Malenka. Ethology of sleep studied with time‐lapse photography: postural immobility and sleep‐cycle phase in humans. Science 201: 1251–1253, 1978.
 322. Hoffer, B. J., G. R. Siggins, A. P. Oliver, and F. E. Bloom. Activation of the pathway from locus coeruleus to rat cerebellar Purkinje neurons: pharmacological evidence of noradrenergic central inhibition. J. Pharmacol. Exp. Ther. 184: 553–569, 1973.
 323. Hoffer, B. J., G. R. Siggins, D. J. Woodward, and F. E. Bloom. Spontaneous discharge of Purkinje neurons after destruction of catecholamine‐containing afferents by 6‐hydroxydopamine. Brain Res. 30: 425–430, 1971.
 324. Hoffman, J. S., and E. Domino. Comparative effects of reserpine on the sleep cycle of man and cat. J. Pharmacol. Exp. Ther. 170: 190–198, 1969.
 325. Holcombe, V., and W. C. Hall. The laminar origin and distribution of the crossed tectoreticular pathways. J. Neurosci. 1: 1103–1112, 1981.
 326. Horvath, F. E., and P. Buser. Thalamo‐caudate‐cortical relationships in synchronized activity. I. Differentiation between ventral and dorsal spindle systems. Brain Res. 39: 21–41, 1972.
 327. Hoshino, K., and O. Pompeiano. Selective discharge of pontine neurons during the postural atonia produced by an anticholinesterase in the decerebrate cat. Arch. Ital. Biol. 114: 244–247, 1976.
 328. Hosoya, Y., and M. Matsushita. Brainstem projections from the lateral hypothalamic area in the rat, as studied with autoradiography. Neurosci. Lett. 24: 111–116, 1981.
 329. Houser, C. R., J. E. Vaughn, R. P. Barber, and E. Roberts. GABA neurons are the major cell type of the nucleus reticularis thalami. Brain Res. 200: 341–354, 1980.
 330. Hubbard, J. E., and V. DiCarlo. Fluorescence histochemistry of monoamine‐containing cell bodies in the brain stem of the squirrel monkey (Saimiri sciureus). J. Comp. Neurol. 147: 553–566, 1973.
 331. Hubel, D. H. Single unit activity in striate cortex of unrestrained cats. J. Physiol. London 147: 226–238, 1959.
 332. Hubel, D. H. Single unit activity in lateral geniculate body and optic tract of unrestrained cats. J. Physiol. London 150: 91–104, 1960.
 333. Hugelin, A. Does the respiratory rhythm originate from a reticular oscillator in the waking state? In: The Reticular Formation Revisited, edited by J. A. Hobson and M. A. B. Brazier. New York: Raven, 1979, p. 261–274.
 334. Huttenlocher, P. R. Evoked and spontaneous activity in single units of medial brain stem during natural sleep and waking. J. Neurophysiol. 24: 451–468, 1961.
 335. Ibuka, N., S. T. Inouye, and H. Hawamura. Analysis of sleep‐wakefulness rhythms in male rats after suprachiasmatic nucleus lesions and ocular enucleations. Brain Res. 122: 33–47, 1977.
 336. Inouye, S. T., and H. Kawamura. Persistence of circadian rhythmicity in a mammalian hypothalamic “island” containing the suprachiasmatic nucleus. Proc. Natl. Acad. Sci. USA 76: 5962–5966, 1979.
 337. Inubushi, S., T. Kobayashi, T. Oshima, and S. Torii. Intracellular recordings from the motor cortex during EEG arousal in unanesthetized brain preparations of the cat. Jpn. J. Physiol. 28: 669–688, 1978.
 338. Inubushi, S., T. Kobayashi, T. Oshima, and S. Torii. An intracellular analysis of EEG arousal in cat motor cortex. Jpn. J. Physiol. 28: 689–708, 1978.
 339. Iskander, T. N., and R. Kaebling. Catecholamine, a dream sleep model and depression. Am. J. Psychiatry 127: 43–50, 1970.
 340. Ito, M., M. Udo, and N. Mano. Long inhibitory and excitatory pathways converging onto cat reticular and Deiters' neurons and their relevance to reticulofugal axons. J. Neurophysiol. 33: 210–226, 1970.
 341. Itoh, K., and N. Mizuno. Direct projections from the mesodiencephalic midline areas to the pericruciate cortex in the cat: an experimental study with the horseradish peroxidase method. Brain Res. 116: 492–497, 1976.
 342. Itoh, K., and N. Mizuno. Topographical arrangement of thalamocortical neurons in the centrolateral nucleus (CL) of the cat, with special reference to a spino‐thalamo‐motor cortical path through the CL. Exp. Brain Res. 30: 471–480, 1977.
 343. Jacobs, B. L., R. Asher, S. J. Henriksen, and W. C. Dement. Electroencephalographic and behavioural effects of stimulation of the raphe nuclei in cats (Abstract). Sleep Res. 1: 23, 1972.
 344. Jacobs, B. L., and B. E. Jones. The role of central monoamine and acetylcholine systems in sleep‐wakefulness states: mediation or modulation. In: Cholinergic‐Monoaminergic Interactions in the Brain, edited by L. L. Butcher. New York: Academic, 1978, p. 271–290.
 345. Jahnsen, H., and R. Llinás. Electrophysiological properties of guinea‐pig thalamic neurones: an in vitro study. J. Physiol. London 349: 205–226, 1984.
 346. Jahnsen, H., and R. Llinás. Ionic basis for the electroresponsiveness and oscillatory properties of guinea‐pig thalamic neurones in vitro. J. Physiol. London 349: 227–247, 1984.
 347. Jalfre, M., M. A. Monachon, and W. Haefely. Drug and PGO‐waves in the cat. In: Proc. Can. Int. Symp. Sleep, 1st, 1972, p. 155–184.
 348. Jalfre, M., M. A. Ruch‐Monachon, and W. Haefely. Methods for assessing the interaction of agents with 5‐hydroxytryptamine neurons and receptors in the brain. Adv. Biochem. Psychopharmacol. 10: 121–134, 1974.
 349. Jankowska, E., S. Lund, A. Lundberg, and O. Pompeiano. Inhibitory effects evoked through ventral reticulospinal pathways. Arch. Ital. Biol. 106: 124–140, 1968.
 350. Jasper, H. H. Diffuse projection systems: the integrative action of the thalamic reticular system. Electroencephalogr. Clin. Neurophysiol. 1: 405–420, 1949.
 351. Jasper, H. H. Recent advances in our understanding of ascending activities of the reticular system. In: Reticular Formation of the Brain, edited by H. H. Jasper, L. D. Proctor, R. S. Knighton, W. C. Noshay, and R. T. Costello. Boston, MA: Little, Brown, 1958, p. 319–331.
 352. Jasper, H. H. Unspecific thalamocortical relations. In: Handbook of Physiology. Neurophysiology, edited by J. Field and H. W. Magoun. Washington, DC: Am. Physiol. Soc., 1960, sect. 1, vol. II, chapt. LIII, p. 1307–1321.
 353. Jasper, H. H. Chairman's summary. In: Corticothalamic Projections and Sensorimotor Activities, edited by T. Frigyesi, E. Rinvik, and M. D. Yahr. New York: Raven, 1972, p. 305–308.
 354. Jasper, H. H. Problems of relating cellular or modular specificity to cognitive functions: importance of state‐dependent reaction. In: The Organization of the Cerebral Cortex, edited by F. O. Schmitt, F. G. Worden, G. Adelman, and S. G. Dennis. Cambridge, MA: MIT Press, 1981, p. 375–393.
 355. Jasper, H. H., and I. Koyama. Rate of release of aminoacids from the cerebral cortex in the cat as affected by brainstem and thalamic stimulation. Can. J. Physiol. Pharmacol. 47: 889–905, 1969.
 356. Jasper, H., R. Naquet, and E. E. King. Thalamocortical recruiting responses in sensory receiving areas in the cat. Electroencephalogr. Clin. Neurophysiol. 7: 99–114, 1955.
 357. Jasper, H., G. F. Rici, and B. Doane. Patterns of cortical neuron discharge during conditioned responses in monkeys. In: Neurological Basis of Behaviour, edited by G. E. W. Wolstenholme and M. J. O'Connor. Boston, MA: Little, Brown, 1958, p. 277–294. (Ciba Found. Symp.)
 358. Jasper, H. H., and J. Tessier. Acetylcholine liberation from cerebral cortex during paradoxical (REM) sleep. Science 172: 601–602, 1971.
 359. Jeannerod, M., and S. Kiyono. Décharge unitaire de la formation réticulée pontique et activité phasique ponto‐géniculo‐occipitale chez le chat sous réserpine. Brain Res. 12: 112–128, 1969.
 360. Jewett, R. E., and S. Norton. Effects of some stimulant and depressant drugs on sleep cycles of cats. Exp. Neurol. 15: 463–474, 1966.
 361. Jinnai, K., and Y. Matsuda. Thalamocaudate projection neurons with a branching axon to the cerebral motor cortex. Neurosci. Lett. 26: 95–99, 1981.
 362. Jones, B. E. The respective involvement of noradrenaline and its deaminated metabolites in waking and paradoxical sleep: a neuropharmacological model. Brain Res. 39: 121–136, 1972.
 363. Jones, B. E., S. T. Harper, and A. E. Halaris. Effects of locus coeruleus lesions upon cerebral monoamine content, sleep‐wakefulness states and the response to amphetamine in the cat. Brain Res. 124: 473–496, 1977.
 364. Jones, B. E., and R. Y. Moore. Catecholamine‐containing neurons of the nucleus locus coeruleus in the cat. J. Comp. Neurol. 157: 42–51, 1974.
 365. Jones, B. E., and R. Y. Moore. Ascending projections of the locus coeruleus in the rat. II. Autoradiographic study. Brain Res. 127: 23–53, 1977.
 366. Jones, E. G. Possible determinants of the degree of retrograde neuronal labeling with horseradish peroxidase. Brain Res. 85: 249–253, 1975.
 367. Jones, E. G. Some aspects of the organization of the thalamic reticular complex. J. Comp. Neurol. 162: 285–308, 1975.
 368. Jones, E. G. Varieties and distribution of non‐pyramidal cells in the somatic sensory cortex of the squirrel monkey. J. Comp. Neurol. 160: 205–267, 1975.
 369. Jones, E. G. Anatomy of cerebral cortex: columnar input‐ouput organization. In: The Organization of the Cerebral Cortex, edited by F. O. Schmitt, F. G. Worden, G. Adelman, and S. G. Dennis. Cambridge, MA: MIT Press, 1981, p. 199–235.
 370. Jones, E. G. Functional subdivision and synaptic organization of the mammalian thalamus. In: Neurophysiology IV, edited by R. Porter. Baltimore, MD: University Park, 1981, vol. 25, p. 173–245. (Int. Rev. Physiol. Ser.)
 371. Jones, E. G., and H. Burton. Cytoarchitecture and somatic sensory connectivity of thalamic nuclei other than the ventrobasal complex in the cat. J. Comp. Neurol. 154: 395–432, 1974.
 372. Jones, E. G., H. Burton, C. B. Saper, and L. W. Swanson. Midbrain, diencephalic and cortical relationships of the basal nucleus of Meynert and associated structures in primates. J. Comp. Neurol. 167: 385–420, 1976.
 373. Jones, E. G., and R. Y. Leavitt. Retrograde axonal transport and the demonstration of nonspecific projections to the cerebral cortex and striatum from thalamic intralaminar nuclei in the rat, cat and monkey. J. Comp. Neurol. 154: 349–378, 1974.
 374. Jouvet, M. Recherches sur les structures nerveuses et les mécanismes responsables des différentes phases du sommeil physiologique. Arch. Ital. Biol. 100: 125–206, 1962.
 375. Jouvet, M. Mechanisms of the states of sleep: a neuropharmacological approach. In: Sleep and Altered States of Consciousness, edited by S. S. Kety. Baltimore, MD: Williams & Wilkins, 1967, p. 86–126.
 376. Jouvet, M. Neurophysiology of the states of sleep. In: The Neurosciences: A Study Program, edited by G. C. Quarton, T. Melnechuk, and F. O. Schmitt. New York: Rockefeller Univ. Press, 1967, p. 529–544.
 377. Jouvet, M. Biogenic amines and the states of sleep. Science 163: 32–41, 1969.
 378. Jouvet, M. The role of monoamines and acetylcholine‐containing neurons in the regulation of the sleep‐waking cycle. Ergeb. Physiol. Biol. Chem. Exp. Pharmakol. 64: 166–307, 1972.
 379. Jouvet, M. Essai sur le réve. Arch. Ital. Biol. 111: 564–576, 1973.
 380. Jouvet, M., and J. F. Delorme. Locus coeruleus et sommeil paradoxal. C. R. Soc. Biol. 159: 895–899, 1965.
 381. Jouvet, M., M. Jeannerod, and F. Delorme. Organisation du système responsable de l'activité phasique au cours du sommeil paradoxal. C. R. Soc. Biol. 7: 1599–1604, 1965.
 382. Jouvet, M., and F. Michel. Correlations électromyographiques du sommeil chez le chat décortiqué et mésencéphalique chronique. C. R. Soc. Biol. 153: 422–425, 1959.
 383. Kaelber, W. W., and T. B. Smith. Projections of the zona incerta in the cat, with stimulation controls. Exp. Neurol. 63: 177–200, 1979.
 384. Kales, A., R. Cadieux, C. R. Soldatos, and T. L. Tan. Successful treatment of narcolepsy with propanolol: a case report. Arch. Neurol. 36: 650–651, 1979.
 385. Kales, A., A. Jacobson, M. J. Paulson, J. Kales, and R. D. Walter. Somnambulism: psychophysiological correlates. I. All‐night EEG studies. Arch. Gen. Psychiatry 14: 586–594, 1966.
 386. Kales, A., and J. D. Kales. Recent findings in the diagnosis and treatment of disturbed sleep. N. Engl. J. Med. 290: 487–499, 1974.
 387. Kales, A., C. R. Solidatos, R. Cadieux, E. O. Bixler, T. L. Tan, and M. B. Scharf. Propranolol in the treatment of narcolepsy. Ann. Intern. Med. 91: 741–743, 1979.
 388. Kanamori, N., K. Sakai, and M. Jouvet. Neuronal activity specific to paradoxical sleep in the ventromedial medullary reticular formation of unrestrained cats. Brain Res. 189: 251–255, 1980.
 389. Kandel, E. R. Cellular Basis of Behavior: An Introduction to Behavioral Neurobiology. San Francisco, CA: Freeman, 1976.
 390. Kasamatsu, T. Maintained and evoked unit activity in the mesencephalic reticular formation of the freely behaving cat. Exp. Neurol. 28: 450–470, 1970.
 391. Kasamatsu, T., and P. Heggelund. Single cell responses in cat visual cortex to visual stimulation during iontophoresis of noradrenaline. Exp. Brain Res. 45: 317–327, 1982.
 392. Kasamatsu, T., and J. D. Pettigrew. Depletion of brain catecholamines: failure of ocular dominance shift after monocular occlusion in kittens. Science 194: 206–208, 1976.
 393. Katayama, Y., Y. Ueno, T. Tsukiyama, and T. Tsubokawa. Long lasting suppression of firing of cortical neurons and decrease in cortical blood flow following train pulse stimulation of the locus coeruleus in the cat. Brain Res. 216: 173–179, 1981.
 394. Keller, E. L. Participation of medial pontine reticular formation in eye movement generation in monkey. J. Neurophysiol. 37: 316–332, 1974.
 395. Kennedy, H., and C. Balaydier. Direct projections from thalamic intralaminar nuclei to extra‐striate visual cortex in the cat traced with horseradish peroxidase. Exp. Brain Res. 28: 133–139, 1977.
 396. Kernell, D., and B. Zwaagstra. Input conductance, axonal conduction velocity and cell size among hindlimb motoneurones of the cat. Brain Res. 204: 311–326, 1981.
 397. Kevetter, G. A., and W. D. Willis. Spinothalamic cells in the rat lumbar cord with collaterals to the medullary reticular formation. Brain Res. 238: 181–185, 1982.
 398. Khazan, N., and P. Brown. Differential effects of three tricyclic antidepressants on sleep and REM sleep in the rat. Life Sci. 9: 279–284, 1970.
 399. Khazan, N., and C. H. Sawyer. Mechanisms of paradoxical sleep as revealed by neurophysiology and pharmacologic approaches in the rabbit. Psychopharmacologia 5: 457–462, 1964.
 400. Kievit, J., and H. G. J. M. Kuypers. Fastigial cerebellar projections to the ventrolateral nucleus of the thalamus and the organization of the descending pathways. In: Corticothalamic Projections and Sensorimotor Activities, edited by T. Frigyesi, E. Rinvik, and M. D. Yahr. New York: Raven, 1972, p. 91–111.
 401. Kievet, J., and H. G. J. M. Kuypers. Basal forebrain and hypothalamic connections to the frontal and parietal cortex in the rhesus monkey. Science 187: 660–662, 1975.
 402. Killackey, H. P., and F. F. Ebner. Two different types of thalamocortical projections to a single cortical area in mammals. Brain Behav. Evol. 6: 141–169, 1972.
 403. Kimura, H., P. L. McGeer, J. H. Peng, and E. G. McGeer. The central cholinergic system studied by choline acetyltransferase immunihistochemistry in the cat. J. Comp. Neurol. 200: 151–201, 1981.
 404. King, A. B., and S. M. Robinson. Ventilation response to hypoxia and acute mountain sickness. Aerosp. Med. 43: 419–421, 1972.
 405. King, C. D. The pharmacology of rapid eye movement sleep. Adv. Pharmacol. Chemother. 9: 1–91, 1971.
 406. King, C. D., and E. J. Jewett. The effects of γ‐methyltyrosine on sleep and brain norepinephrine in cats. J. Pharmacol. Exp. Ther. 177: 188–195, 1971.
 407. Kitsikis, A., and M. Steriade. Thalamic, callosal and reticular converging inputs to parietal association cortex in cat. Brain Res. 93: 516–524, 1975.
 408. Kitsikis, A., and M. Steriade. Immediate behavioral effects of kainic acid injections into the midbrain reticular core. Behav. Brain Res. 3: 361–380, 1981.
 409. Klee, M. R., H. D. Lux, and K. Offenloch. Veränderungen der Membranpolarisation und der Erregbarkeit von Zellen der motorischen Rinde während hochfrequenter Reizung der Formation Reticularis. Arch. Psychol. Gesamte Neurol. 205: 237–261, 1974.
 410. Kleitman, N. Sleep and Wakefulness. Chicago, IL: Univ. of Chicago Press, 1963.
 411. Knight, D. P. Histochemical demonstration of catecholamines and acetylcholinesterase in the same cell bodies in the locus coeruleus (rat hindbrain). Proc. R. Microsc. Soc. 6: 26–27, 1970.
 412. Kornhuber, H. H., and L. Deecke. Hirnpotentialànderungen bei Willkürbewegungen und passiven Bewegunger des Menschen: Bereitschaftspotential und reafferente Potentiale. Pfluegers Arch. Gesamte Physiol. Menschen Tiere 284: 1–17, 1965.
 413. Kostowski, W., E. Giacalone, S. Garattini, and L. Valzelli. Electrical stimulation of midbrain raphe: biochemical, behavioural and bioelectrical effects. Eur. J. Pharmacol. 7: 170–175, 1969.
 414. Krammer, E. B., M. F. Lischka, M. Karobath, and G. Schonbeck. Is there a selectivity of neuronal degeneration induced by intrastriatal injection of kainic acid? Brain Res. 177: 577–582, 1979.
 415. Kreindler, A., and M. Steriade. EEG patterns of arousal and sleep induced by stimulating various amygdaloid levels in the cat. Arch. Ital. Biol. 102: 576–586, 1964.
 416. Krettek, J. E., and J. L. Price. The cortical projections of the mediodorsal nucleus and adjacent thalamic nuclei in the rat. J. Comp. Neurol. 171: 157–192, 1977.
 417. Krnjević, K. Chemical nature of synaptic transmission in vertebrates. Physiol. Rev. 54: 418–540, 1974.
 418. Krnjević, K. Transmitters in motor systems. In: Handbook of Physiology. The Nervous System. Motor Control, edited by J. M. Brookhart and V. B. Mountcastle. Bethesda, MD: Am. Physiol. Soc., 1981, sect. 1, vol. II, pt. 1, chapt. 4, p. 107–154.
 419. Krnjević, K., and J. W. Phillis. Acetylcholine‐sensitive cells in the cerebral cortex. J. Physiol. London 166: 296–327, 1963.
 420. Krnjević, K., E. Puil, and R. Werman. EGTA and motoneuronal after‐potentials. J. Physiol. London 275: 199–223, 1978.
 421. Krnjević, K., R. Pumain, and L. Renaud. The mechanism of excitation by acetylcholine in the cerebral cortex. J. Physiol. London 215: 247–268, 1971.
 422. Krnjević, K., and W. Reinhardt. Choline excites cortical neurons. Science 206: 1321–1323, 1979.
 423. Krnjević, K., and S. Schwartz. Some properties of unresponsive cells in the cerebral cortex. Exp. Brain Res. 3: 306–319, 1967.
 424. Kromer, L. F., and R. Y. Moore. A study of the organization of the locus coeruleus projections to the lateral geniculate nuclei in the albino rat. Neuroscience 5: 255–271, 1980.
 425. Kucera, P., and P. Favrod. Suprachiasmatic nucleus projection to mesencephalic central gray in the woodmouse (Apodemus sylvaticus L.). Neuroscience 4: 1705–1715, 1979.
 426. Kuhar, M. J., G. K. Aghajanian, and R. H. Roth. Tryptophan hydroxylase activity and synaptosomal uptake of serotonin in discrete brain regions after midbrain raphe lesions: correlations with serotonin levels and histochemical fluorescence. Brain Res. 44: 165–176, 1972.
 427. Kunze, W., J. S. McKenzie, and A. P. Bendrups. An electrophysiological study of thalamo‐caudate neurones in the cat. Exp. Brain Res. 36: 233–244, 1979.
 428. Kupfer, D. J., and M. B. Bowers Jr. REM sleep and central monoamine oxidase inhibition. Psychopharmacologia 27: 183–190, 1972.
 429. Kurtz, D., J. Krieger, and J. C. Stierle. EMG activity of cricothyroid and chin muscles during wakefulness and sleeping in the sleep apnea syndrome. Electroencephalogr. Clin. Neurophysiol. 45: 777–784, 1978.
 430. Kuypers, H. G. J. M. Cortico‐bulbar connexcions to the pons and lower brain stem in man. An anatomical study. Brain 81: 364–388, 1958.
 431. Kuypers, H. G. J. M., and V. A. Maisky. Retrograde axonal transport of horseradish peroxidase from spinal cord to brain stem cell group in the cat. Neurosci. Lett. 1: 9–14, 1975.
 432. Laffont, F., A. Autret, M. Minz, T. Beillevaire, H. P. Cathela, and P. Castaigne. Sleep respiratory arrhythmias in control subjects, narcoleptics and non‐cataplectic hypersomniacs. Electroencephalogr. Clin. Neurophysiol. 44: 697–705, 1978.
 433. Lamarre, Y., M. Filion, and J. P. Cordeau. Neuronal discharges of the ventrolateral nucleus of the thalamus during sleep and wakefulness in the cat. I. Spontaneous activity. Exp. Brain Res. 12: 480–498, 1971.
 434. Lamour, Y., P. Dutar, and A. Jobert. Excitatory effect of acetylcholine on different types of neurons in the first somatosensory neocortex of the rat: laminar distribution and pharmacological characteristics. Neuroscience 7: 1483–1494, 1982.
 435. Laurent, J. P., and F. Ayalaguerrero. Reversible suppression of ponto‐geniculo‐occipital waves by localized cooling during paradoxical sleep in cat. Exp. Neurol. 49: 356–369, 1975.
 436. Laurent, J. P., R. Cespuglio, and M. Jouvet. Delimitation des voies ascendantes de l'activité ponto‐geniculo‐occipitale chez le chat. Brain Res. 65: 29–52, 1974.
 437. Leger, L., L. Wiklund, L. Descarries, and M. Persson. Description of an indolaminergic cell component in the cat locus coeruleus: a fluorescence histochemical and radioautographic study. Brain Res. 168: 43–56, 1979.
 438. Le Gros Clark, W. E. The structure and connections of the thalamus. Brain 55: 406–470, 1932.
 439. Lehmann, J., J. I. Nagy, S. Atmadja, and H. C. Fibiger. The nucleus basalis magnocellularis: the origin of a cholinergic projection to the neocortex of the rat. Neuroscience 5: 1161–1174, 1980.
 440. Leontovitch, T. A., and G. P. Zhukova. The specificity of the neuronal structure and topography of the reticular formation in the brain and spinal cord of carnivora. J. Comp. Neurol. 121: 347–389, 1963.
 441. Libet, B., H. Kobayashi, and T. Tanaka. Synaptic coupling into the production and storage of a neuronal memory trace. Nature London 258: 155–157, 1975.
 442. Lidbrink, P., G. Jonsson, and K. Fuxe. The effect of imipramine‐like drugs and antihistamine drugs on uptake mechanisms in the central noradrenaline and 5‐hydroxytryptamine neurons. Neuropharmacology 10: 521–536, 1971.
 443. Lidov, H. G. W., R. Grzanna, and M. E. Molliver. The serotonin innervation of the cerebral cortex in the rat—an immunohistochemical analysis. Neuroscience 5: 207–227, 1980.
 444. Lindsley, D. B., J. W. Bowden, and H. W. Magoun. Effect upon the EEG of acute injury to the brain stem activating system. Electroencephalogr. Clin. Neurophysiol. 1: 475–486, 1949.
 445. Lindsley, D. B., L. H. Schreiner, W. B. Knowles, and H. W. Magoun. Behavioral and EEG changes following chronic brain stem lesions in the cat. Electroencephalogr. Clin. Neurophysiol. 2: 483–498, 1950.
 446. Lindstrom, S. Synaptic organization of inhibitory pathways to principal cells in the lateral geniculate nucleus of the cat. Brain Res. 234: 447–453, 1982.
 447. Lindvall, O., and A. Björklund. The organization of the ascending catecholamine neuron systems in the rat brain as revealed by the glyoxylic acid fluorescence method. Acta Physiol. Scand. Suppl. 412: 1–48, 1974.
 448. Lindvall, O., A. Björklund, R. Y. Moore, and U. Stenevi. Mesencephalic dopamine neurons projecting to neocortex. Brain Res. 81: 325–331, 1974.
 449. Lindvall, O., A. Björklund, A. Nobin, and U. Stenevi. The adrenergic innervation of the rat thalamus as revealed by the glyoxylic acid fluorescence method. J. Comp. Neurol. 154: 317–348, 1974.
 450. Lipski, J. Antidromic activation of neurones as an analytical tool in the study of the central nervous system. J. Neurosci. Methods 4: 1–32, 1981.
 451. Livingston, M. S., and D. H. Hubel. Effects of sleep and arousal on the processing of visual information in the cat. Nature London 291: 554–561, 1981.
 452. Ljungberg, T., and U. Ungerstedt. Sensory inattention produced by 6‐hydroxydopamine‐induced degeneration of ascending dopamine neurons in the brain. Exp. Neurol. 53: 585–600, 1976.
 453. Llinás, R., and H. Jahnsen. Electrophysiology of mammalian thalamic neurones in vitro. Nature London 297: 406–408, 1982.
 454. Llinás, R., and Y. Yarom. Electrophysiology of mammalian inferior olivary neurones in vitro. Different types of voltage‐dependent ionic conductance. J. Physiol. London 315: 549–567, 1981.
 455. Llinás, R., and Y. Yarom. Properties and distribution of ionic conductances generating electroresponsiveness of mammalian inferior olivary neurones in vitro. J. Physiol. London 315: 569–584, 1981.
 456. Lorente De Nó, R. Cerebral cortex: architecture, intracortical connections, motor projections. In: Physiology of the Nervous System, edited by J. F. Fulton. New York: Oxford Univ. Press, 1938, p. 291–340.
 457. Lotka, A. Elements of Physical Biology. New York: Dover, 1956.
 458. Lucero, M. A. Lengthening of REM sleep duration consecutive to learning in the rat. Brain Res. 20: 319–322, 1970.
 459. Lynch, J. C., V. B. Mountcastle, W. H. Talbot, and T. C. T. Yin. Parietal lobe mechanisms for directed visual attention. J. Neurophysiol. 40: 362–389, 1977.
 460. Macchi, G., M. Bentivoglio, C. D'Atena, P. Rossini, and E. Tempesta. The cortical projection of the thalamic intra‐laminar nuclei restudied by means of the HRP retrograde axonal transport. Neurosci. Lett. 4: 121–126, 1977.
 461. Macchi, G., M. Bentivoglio, M. Molinari, and D. Minciacchi. The thalamo‐caudate versus thalamo‐cortical projections as studied in the cat with fluorescent retrograde double labeling. Exp. Brain Res. 54: 225–239, 1984.
 462. MacGregor, R. J., R. Prieto‐Diaz, S. W. Miller, and P. M. Groves. Statistical properties of neurons in the rat mesencephalic reticular formation. Brain Res. 64: 167–187, 1973.
 463. Madison, D. V., and R. A. Nicoll. Noradrenaline blocks accommodation of pyramidal cell discharge in the hippocampus. Nature London 299: 636–638, 1982.
 464. Maeda, T., C. Pin, D. Salvert, L. Leger, and M. Jouvet. Les neurones contenant des catecholamines du tegmentum pontique et leurs voies de projection chez le chat. Brain Res. 57: 119–152, 1973.
 465. Maekawa, K., and D. P. Purpura. Properties of spontaneous and evoked synaptic activities of thalamic ventrobasal neurons. J. Neurophysiol. 30: 360–381, 1967.
 466. Magni, F., and W. D. Willis. Identification of reticular formation neurons by intracellular recording. Arch. Ital. Biol. 101: 681–702, 1963.
 467. Magni, F., and W. D. Willis. Cortical control of brain stem reticular neurons. Arch. Ital. Biol. 102: 418–433, 1964.
 468. Magni, F., and W. D. Willis. Subcortical and peripheral control of brain stem reticular neurons. Arch. Ital. Biol. 102: 434–448, 1964.
 469. Magoun, H. W., and R. Rhines. An inhibitory mechanism in the bulbar reticular formation. J. Neurophysiol. 9: 165–171, 1946.
 470. Malthe‐Sørenssen, D., E. Odden, and I. Walaas. Selective destruction by kainic acid of neurons innervated by putative glutamergic afferents in septum and nucleus of the diagonal band. Brain Res. 182: 461–465, 1980.
 471. Mancia, M. Electrophysiological and behavioural changes owing to splitting of the brain stem in cat. Electroencephalogr. Clin. Neurophysiol. 27: 487–502, 1969.
 472. Mancia, M., A. Grantyn, G. Broggi, and M. Margnelli. Synaptic linkage between mesencephalic and bulbo‐pontine reticular structures as revealed by intracellular recording. Brain Res. 33: 491–494, 1971.
 473. Mancia, M., M. Margnelli, M. Mariotti, R. Spreafico, and G. Broggi. Brain stem‐thalamus reciprocal influences in the cat. Brain Res. 69: 297–314, 1974.
 474. Mancia, M., M. Mariotti, E. R. Roman, and M. Schipatti. Basal forebrain and hypothalamic influences upon brain stem neurons. Brain Res. 107: 487–497, 1976.
 475. Mannen, H. “Noyau fermé” et “noyau ouvert”. Contribution à l'étude cytoarchitectonique du tronc cérébral envisagée du point de vue du mode d'arborisation dendritique. Arch. Ital. Biol. 98: 333–350, 1960.
 476. Manohar, S., H. Noda, and W. R. Adey. Behavior of mesencephalic reticular neurons in sleep and wakefulness. Exp. Neurol. 34: 140–157, 1972.
 477. Mantegazzini, P., and G. Pepeu. Increase of cortical acetylcholine induced by midbrain hemisection in the cat. J. Physiol. London 173: 20, 1964.
 478. Markowitsch, H. J., and E. Irle. Widespread cortical projections of the ventral tegmental area and of other brain stem structures in the cat. Exp. Brain Res. 41: 233–246, 1981.
 479. Marshall, J. F., S. Richardson, and P. Teitelbaum. Nigrostriatal bundle damage and the lateral hypothalamic syndrome. J. Comp. Physiol. Psychol. 87: 808–830, 1974.
 480. Marshall, J. F., B. M. Turner, and P. Teitelbaum. Sensory neglect produced by lateral hypothalamic damage. Science 174: 523–525, 1971.
 481. Marshall, K. C., and H. McLennan. The synaptic activation of neurones of the feline ventrolateral thalamic nucleus: possible cholinergic mechanisms. Exp. Brain Res. 15: 472–483, 1972.
 482. Marshall, K. C., and J. S. Murray. Cholinergic facilitation of thalamic relay transmission in the cat. Exp. Neurol. 69: 318–333, 1980.
 483. Mason, S. T., and H. C. Fibiger. Regional topography within noradrenergic locus coeruleus as revealed by retrograde transport of horseradish peroxidase. J. Comp. Neurol. 187: 703–724, 1979.
 484. Mason, S. T., and H. C. Fibiger. On the specificity of kainic acid. Science 204: 1339–1341, 1979.
 485. Mason, S. T., and S. D. Iversen. Learning in the absence of forebrain noradrenaline. Nature London 258: 422–424, 1975.
 486. Masserano, J. M., and C. King. Effects on sleep of acetylcholine perfusion of the locus coeruleus of cats. Neuropharmacology 21: 1163–1167, 1982.
 487. Matsumoto, J., and M. Jouvet. Effets de réserpine, dopa et 5‐HTP sur les deux états de sommeil. C. R. Soc. Biol. 158: 2137–2140, 1964.
 488. Matsuzaki, M. Differential effects of Na‐butyrate and physostigmine on brain stem activities of sleep. Brain Res. 11: 251–255, 1968.
 489. Maunz, R. A., N. G. Pitts, and B. W. Peterson. Cat spinoreticular neurons: locations, responses and changes in responses during repetitive stimulation. Brain Res. 148: 365–379, 1978.
 490. McBride, R. L., and J. Sutin. Projections of the locus coeruleus and adjacent pontine tegmentum in the cat. J. Comp. Neurol. 165: 265–284, 1976.
 491. McCarley, R. W. Control of sleep‐waking state alteration in Felix domesticus. In: Aspects of Behavioral Neurobiology, edited by J. A. Ferrendelli. Bethesda, MD: Soc. Neurosci., 1977, vol. 3, p. 90–128. (Soc. Neurosci. Symp.)
 492. McCarley, R. W., O. Benoit, and G. Barrionuevo. Lateral geniculate nucleus unitary discharge in sleep and waking: state‐ and rate‐specific aspects. J. Neurophysiol. 50: 798–818, 1983.
 493. McCarley, R. W., and J. A. Hobson. Single neuron activity in cat gigantocellular tegmental field: selectivity of discharge in desynchronized sleep. Science 174: 1250–1252, 1971.
 494. McCarley, R. W., and J. A. Hobson. Simple spike firing patterns of cat cerebellar Purkinje cells in sleep and waking. Electroencephalogr. Clin. Neurophysiol. 33: 471–483, 1972.
 495. McCarley, R. W., and J. A. Hobson. Discharge patterns of cat pontine brain stem neurons during desynchronized sleep. J. Neurophysiol. 38: 751–766, 1975.
 496. McCarley, R. W., and J. A. Hobson. Neuronal excitability modulation over the sleep cycle: a structural and mathematical model. Science 189: 58–60, 1975.
 497. McCarley, R. W., and J. A. Hobson. Mind‐body isomorphism and the study of dreams. In: Sleep, Dreams and Memory, edited by W. Fishbein. New York: Spectrum, 1981.
 498. McCarley, R. W., and K. Ito. The pontine reticular formation: an intracellular recording study of inputs from midbrain, bulbar and contralateral pontine reticular formation in the unanesthetized cat. Soc. Neurosci. Abstr. 6: 203, 1980.
 499. McCarley, R. W., and K. Ito. Membrane potential depolarization of PRF neurons during behavioral state changes from synchronized to desynchronized sleep in naturally sleeping cats. Soc. Neurosci. Abstr. 8: 726, 1982.
 500. McCarley, R. W., J. P. Nelson, and J. A. Hobson. Ponto‐geniculo‐occipital (PGO) burst neurons: correlative evidence for neuronal generators of PGO waves. Science 201: 269–272, 1978.
 501. McGinty, D. J., and R. M. Harper. Dorsal raphe neurons: depression of firing during sleep in cats. Brain Res. 101: 569–575, 1976.
 502. McGinty, D. J., R. M. Harper, and M. K. Fairbanks. 5‐HT‐containing neurons: unit activity in behaving cats. In: Serotonin and Behavior, edited by J. Barchas and E. Usdin. New York: Academic, 1973, p. 267–269.
 503. McGinty, D. J., M. S. London, T. L. Baker, M. Stevenson, T. Hoppenbrouwers, R. M. Harper, M. B. Sterman, and J. Hodgman. Sleep apnea in normal kittens. Sleep 1: 393–412, 1979.
 504. McGinty, D. J., and J. M. Siegel. Sleep states. In: Handbook of Behavioral Neurobiology. Motivation, edited by E. Satinoff and P. Teitelbaum. New York: Plenum, 1983, vol. 6, p. 105–181.
 505. McGuiness, C. M., and G. M. Krauthamer. The afferent projections to the centrum medianum of the cat as demonstrated by retrograde transport of horseradish peroxidase. Brain Res. 184: 255–269, 1980.
 506. McKenna, T., R. W. McCarley, T. Amatruda, D. Black, and J. A. Hobson. Effects of carbachol at pontine sites yielding long duration desynchronized sleep episodes (Abstract). Sleep Res. 3: 39, 1974.
 507. McMahon, D. Chemical messengers in development: a hypothesis. Science 185: 1012–1021, 1974.
 508. Meech, R. W. The sensitivity of Helix aspersa neurones to injected calcium ions. J. Physiol. London 237: 259–277, 1974.
 509. Mehler, W. R. Further notes on the center median nucleus of Luys. In: The Thalamus, edited by D. P. Purpura and M. D. Yahr. New York: Columbia Univ. Press, 1966, p. 109–122.
 510. Menétrey, D., A. Chaouch, and J. M. Besson. Location and properties of dorsal horn neurons at origin of spinoreticular tract in lumbar enlargement of the rat. J. Neurophysiol. 44: 862–877, 1980.
 511. Mikiten, I. H., P. H. Niebyl, and C. D. Hendley. EEG desynchronization during behavioral sleep associated with spike discharges from the thalamus of the cat (Abstract). Federation Proc. 20: 327, 1961.
 512. Miller, A. J., and G. Fry. Postnatal development of presynaptic terminals in the gigantocellular tegmental field (FTG) of the rat. Brain Res. 188: 301–317, 1980.
 513. Minderhoud, J. M. An anatomical study of the efferent connections of the thalamic reticular nucleus. Exp. Brain Res. 12: 435–446, 1971.
 514. Mitler, M. M., and W. C. Dement. Cataplectic‐like behavior in cats after microinjections of carbachol in pontine reticular formation. Brain Res. 68: 335–343, 1964.
 515. Mitler, M. M., and W. C. Dement. Sleep studies on canine narcolepsy: pattern and cycle comparisons between affected and normal dogs. Electroencephalogr. Clin. Neurophysiol. 43: 691–699, 1977.
 516. Mizuno, N., C. D. Clemente, and E. K. Sauerland. Fiber projections from rostral basal forebrain structures in the cat. Exp. Neurol. 25: 220–237, 1969.
 517. Mizuno, N., M. Uemura‐Sumi, Y. Yasui, A. Konishi, and R. Matsushima. Direct projections from the extrathalamic forebrain structures to the neocortex in the macaque monkey. Neurosci. Lett. 29: 13–17, 1982.
 518. Molliver, M. E. Role of monoamines in the development of the neocortex. Neurosci. Res. Program Bull. 20: 492–507, 1982.
 519. Molliver, M. E., and D. A. Kristt. The fine structural demonstration of monoaminergic synapses in immature rat neocortex. Neurosci. Lett. 1: 305–310, 1975.
 520. Monaco, A. P., H. A. Baghdoyan, J. P. Nelson, and J. A. Hobson. Cortical wave amplitude and eye direction movement direction are correlated in REM sleep but not in waking. Arch. Ital. Biol. 122: 213–223, 1984.
 521. Monnier, M., A. M. Hatt, L. B. Cueni, and G. A. Schoenenberger. Humoral transmission of sleep. VI. Purification and assessment of a hypnogenic fraction of “sleep Dialysate” (factor delta). Pfluegers Arch. 331: 257–265, 1972.
 522. Montero, V. M., and G. L. Scott. Synaptic terminals in the dorsal lateral geniculate nucleus from neurons of the thalamic reticular nucleus: a light and electron microscope autoradiographic study. Neuroscience 6: 2561–2577, 1981.
 523. Moore, R. Y. The anatomy of central neural mechanisms regulating endocrine rhythms. In: Endocrine Rhythms, edited by D. T. Krieger. New York: Raven, 1979, p. 63–87.
 524. Moore, R. Y., and F. E. Bloom. Central catecholamine neuron systems: anatomy and physiology of the norepinephrine and epinephrine systems. Annu. Rev. Neurosci. 2: 113–168, 1979.
 525. Moore, R. Y., A. I. Halaris, and B. E. Jones. Serotonin neurons of the midbrain raphe ascending projections. J. Comp. Neurol. 180: 417–438, 1978.
 526. Moore, R. Y., and D. C. Klein. Visual pathways and the central neural control of a circadian rhythm in pineal serotonin N‐acetyltransferase activity. Brain Res. 71: 17–33, 1974.
 527. Morales, F., and M. H. Chase. Postsynaptic control of lumbar motoneuron excitability during active sleep in the chronic cat. Brain Res. 225: 279–295, 1981.
 528. Morin, D., and M. Steriade. Development from primary to augmenting responses in primary somatosensory cortex. Brain Res. 205: 49–66, 1981.
 529. Morison, R. S., and D. L. Bassett. Electrical activity of the thalamus and basal ganglia in decorticate cats. J. Neurophysiol. 8: 309–314, 1945.
 530. Morison, R. S., and E. W. Dempsey. A study of thalamocortical relations. Am. J. Physiol. 135: 281–292, 1942.
 531. Moroni, F., C. Bianchi, G. Moneti, S. Tanganelli, G. Spidelieri, P. Guandalini, and L. Beani. Release of GABA from the guinea‐pig neocortex induced by electrical stimulation of the locus coeruleus or by norepinephrine. Brain Res. 232: 216–221, 1982.
 532. Morrell, J. I., L. M. Greenberger, and D. W. Pfaff. Hypothalamic, other diencephalic and telencephalic neurons that project to the dorsal midbrain. J. Comp. Neurol. 201: 589–620, 1981.
 533. Morrison, A. R., and O. Pompeiano. An analysis of the supraspinal influences acting on motoneurons during sleep in the unrestrained cat. Responses of the alpha motoneurons to direct electrical stimulation during sleep. Arch. Ital. Biol. 103: 497–516, 1965.
 534. Morrison, A. R., and O. Pompeiano. Central depolarization of group Ia afferent fibers during desynchronized sleep. Arch. Ital. Biol. 103: 517–537, 1965.
 535. Morrison, A. R., and O. Pompeiano. Pyramidal discharge from somatosensory cortex and cortical control of primary afferents during sleep. Arch. Ital. Biol. 103: 538–568, 1965.
 536. Morrison, J. H., R. Grzanna, M. E. Molliver, and J. T. Coyle. The distribution and orientation of noradrenergic fibers in neocortex of the rat: an immunofluorescence study. J. Comp. Neurol. 181: 17–40, 1978.
 537. Morrison, J. H., M. E. Molliver, and R. Grzanna. Noradrenergic innervation of cerebral cortex: widespread effects of local cortical lesions. Science 205: 313–316, 1979.
 538. Moruzzi, G. Synchronizing influences of the brain stem and the inhibitory mechanisms underlying the production of sleep by sensory stimulation. Electroencephalogr. Clin. Neurophysiol. Suppl. 13: 231–256, 1960.
 539. Moruzzi, G. The functional significance of sleep with particular regard to the brain mechanisms underlying consciousness. In: Brain and Conscious Experience, edited by J. C. Eccles. New York: Springer‐Verlag, 1966, p. 345–388.
 540. Moruzzi, G. Sleep and instinctive behavior. Arch. Ital. Biol. 107: 175–216, 1969.
 541. Moruzzi, G. The sleep‐waking cycle. Ergeb. Physiol. Biol. Chem. Exp. Pharmakol. 64: 1–165, 1972.
 542. Moruzzi, G., and H. W. Magoun. Brainstem reticular formation and activation of the EEG. Electroencephalogr. Clin. Neurophysiol. 1: 455–473, 1949.
 543. Motter, B. C., and V. B. Mountcastle. The functional properties of the light‐sensitive neurons of the posterior parietal cortex studied in waking monkeys: foveal sparing and opponent vector organization. J. Neurosci. 1: 3–26, 1981.
 544. Mountcastle, V. B., R. A. Andersen, and B. C. Motter. The influence of attentive fixation upon the excitability of the light‐sensitive neurons of the posterior parietal cortex. J. Neurosci. 1: 1218–1235, 1981.
 545. Mountcastle, V. B., J. C. Lynch, A. Georgopoulos, H. Sakata, and C. Acuna. Posterior parietal association cortex of the monkey: command functions for operations within extrapersonal space. J. Neurophysiol. 38: 871–908, 1975.
 546. Mouret, J., and J. Coindet. Polygraphic evidence against a critical role of the raphe nuclei in sleep in the rat. Brain Res. 186: 273–287, 1980.
 547. Mouret, J., J. Coindet, G. Debilly, and G. Chouvet. Suprachiasmatic nuclei lesions in the rat: alterations in sleep circadian rhythms. Electroencephalogr. Clin. Neurophysiol. 45: 402–408, 1978.
 548. Mukhametov, L. M., G. Rizzolatti, and A. Seitun. An analysis of the spontaneous activity of lateral geniculate neurons and of optic tract fibers in free moving cats. Arch. Ital. Biol. 108: 325–347, 1970.
 549. Mukhametov, L. M., G. Rizzolatti, and V. Tradardi. Spontaneous activity of nucleus reticularis thalami in freely moving cats. J. Physiol. London 210: 651–667, 1970.
 550. Nakahama, H., K. Shima, M. Yamamoto, and K. Aya. Regularity of the spontaneous discharge of neurons in the nucleus raphe dorsalis of the cat. Neurosci. Lett. 23: 161–165, 1981.
 551. Nakai, Y., and S. Takaori. Influence of norepinephrine‐containing neurons derived from the locus coeruleus on lateral geniculate neuronal activities of cats. Brain Res. 71: 47–60, 1974.
 552. Nakamura, R. H., C. Kennedy, J. C. Fillin, S. Suda, M. Ito, F. Storch, W. Mendelson, M. Mishkin, and L. Sokoloff. Hypnogenic‐center theory of sleep: no support from 2‐deoxyglucose studies in monkeys. Soc. Neurosci. Abstr. 7: 234, 1981.
 553. Nakamura, S., and K. Iwama. Antidromic activation of the rat locus coeruleus neurons from hippocampus, cerebral and cerebellar cortices. Brain Res. 99: 372–376, 1975.
 554. Nakamura, Y., L. Goldberg, S. Chandler, and M. Chase. Intracellular analysis of trigeminal motoneuron activity during sleep in the cat. Science 199: 204–207, 1978.
 555. Nakamura, Y., and J. Schlag. Cortically induced rhythmic activities in the thalamic ventrolateral nucleus of the cat. Exp. Neurol. 22: 209–221, 1968.
 556. Nakano, K., T. Takimoto, T. Kayahara, Y. Takeuchi, and Y. Kobaysahi. Distribution of cerebellothalamic neurons projecting to the ventral nuclei of the thalamus: an HRP study in the cat. J. Comp. Neurol. 194: 427–439, 1980.
 557. Nakazawa, Y., H. Tachibana, M. Kotorii, and M. Ogata. Effects of L‐dopa on natural night sleep and on rebound of REM sleep. Folia Psychiatr. Neurol. Jpn. 27: 223–230, 1973.
 558. Nauta, W. J. H. Hippocampal projections and related neural pathways to the midbrain in the cat. Brain 81: 319–340, 1958.
 559. Nauta, W. J. H. Projections of the pallidal complex: an autoradiographic study in the cat. Neuroscience 4: 1853–1873, 1979.
 560. Nauta, W. J. H., and M. Cole. Efferent projections of the subthalamic nucleus: an autoradiographic study in monkey and cat. J. Comp. Neurol. 180: 1–16, 1978.
 561. Nauta, W. J. H., and H. G. J. M. Kuypers. Some ascending pathways in the brain stem reticular formation. In: Reticular Formation of the Brain, edited by H. H. Jasper, L. D. Proctor, R. S. Knighton, W. C. Noshay, and R. T. Costello. Boston, MA: Little, Brown, 1958, p. 3–30.
 562. Nauta, W. J. H., and W. R. Mehler. Projections of the lentiform nucleus in the monkey. Brain Res. 1: 3–42, 1966.
 563. Nauta, H. J. W., M. B. Pritz, and R. J. Lassek. Afferents to the rat caudatoputamen studied with horseradish peroxidase: an evaluation of a retrograde neuroanatomical research method. Brain Res. 67: 219–238, 1974.
 564. Nelson, J. P., R. W. McCarley, and J. A. Hobson. REM sleep burst neurons, PGO waves, and eye movement information. J. Neurophysiol. 50: 784–797, 1983.
 565. Netick, A., J. Orem, and W. Dement. Neuronal activity specific to REM sleep and its relationship to breathing. Brain Res. 120: 197–207, 1977.
 566. Nicoll, R. A., and B. E. Alger. Synaptic excitation may activate a calcium‐dependent potassium conductance in hippocampal pyramidal cells. Science 212: 957–959, 1981.
 567. Nieouellon, A., A. Cheramy, and J. Glowinski. Release of dopamine in vivo from cat substantia nigra. Nature London 266: 375–377, 1977.
 568. Niimi, K., S. Kishi, M. Miki, and S. Fujita. An experimental study of the course and termination of the projection fibers from cortical areas 4 and 6 in the cat. Folia Psychiatr. Neurol. Jpn. 17: 167–216, 1963.
 569. Niimi, K., and E. Kuwahara. The dorsal thalamus of the cat and comparison with monkey and man. J. Hirnforsch. 14: 303–325, 1973.
 570. Niimi, K., M. Niimi, and Y. Okada. Thalamic afferents to the limbic cortex in the cat studied with the method of retrograde axonal transport of horseradish peroxidase. Brain Res. 145: 225–238, 1978.
 571. Oakson, G., and M. Steriade. Slow rhythmic rate fluctuations of cat midbrain reticular neurons in synchronized sleep and waking. Brain Res. 247: 277–288, 1982.
 572. Oderfeld‐Nowak, B., A. Potempska, and J. Oderfeld. Analysis of the time course of changes in hippocampal acetylcholinesterase and choline acetyltransferase activities after various septal lesions in the rat: return of enzymic activity after extensive medioventral lesions. Neuroscience 2: 641–648, 1977.
 573. Oertel, W. H., A. M. Graybiel, E. Mugnaini, R. P. Elde, D. E. Schmechel, and I. J. Kopin. Coexistence of glutamic acid decarboxylase and somatostatinlike immunoreactivity in neurons of the feline nucleus reticularis thalami. J. Neurosci. 3: 1322–1332, 1983.
 574. Ohara, P. T., and A. R. Liebermann. Thalamic reticular nucleus: anatomical evidence that cortico‐reticular axons establish monosynaptic contact with reticulo‐geniculate projection cells. Brain Res. 207: 153–156, 1981.
 575. Ohara, P. T., A. J. Sefton, and A. R. Liebermann. Mode of termination of afferents from the thalamic reticular nucleus in the dorsal lateral geniculate nucleus of the rat. Brain Res. 197: 503–506, 1980.
 576. Olivier, A., A. Parent, and L. J. Poirier. Identification of the thalamic nuclei on the basis of their cholinesterase content in the monkey. J. Anat. 106: 37–50, 1970.
 577. Olschowska, J. A., R. Grzanna, and M. E. Molliver. The distribution and incidence of synaptic contacts of noradrenergic varicosities in the rat neocortex: an immunocytochemical study. Soc. Neurosci. Abstr. 6: 352, 1980.
 578. Olson, L., and K. Fuxe. On the projections from the locus coeruleus noradrenaline neurons: the cerebellar innervation. Brain Res. 28: 165–171, 1971.
 579. Olszewski, J., and D. Baxter. The Cytocarchitecture of the Human Brain Stem. New York: Karger, 1954.
 580. Oshtory, M. A., and N. Vijayan. Clorazepam treatment of insomnia due to sleep myoclonus. Arch. Neurol. 37: 119–120. 1980.
 581. Oswald, I. Human brain protein drugs and dreams. Nature London 223: 893–897, 1969.
 582. Oswald, I., K. Adam, S. Allen, R. Burack, M. Spence, and V. Thacore. Alpha adrenergic blocker, thymoxamine and mesoridaxine both increase human REM sleep duration (Abstract). Sleep Res. 3: 62, 1974.
 583. Oswald, I., V. R. Thacore, K. Adam, V. Brezinova, and R. Burack. α‐Adrenergic blockade increases human REM sleep. Br. J. Clin. Pharmacol. 2: 107–110, 1975.
 584. Ottersen, O. P., and Y. Ben‐Ari. Demonstration of a heavy projection of midline thalamic neurons upon the lateral nucleus of the amygdala of the rat. Neurosci. Lett. 9: 147–152, 1978.
 585. Pappenheimer, J. R., G. Koski, V. Fencl, M. L. Karnovsky, and J. Krueger. Extraction of sleep‐promoting factor S from cerebrospinal fluid and from brains of sleep‐deprived animals. J. Neurophysiol. 38: 1299–1311, 1975.
 586. Parent, A. Comparative anatomy of monoaminergic systems. In: Handbook of Chemical Neuroanatomy, edited by A. Björklund, T. Hökfelt, and M. Kuhar. Amsterdam: Elsevier, 1984, vol. 2, p. 409–439.
 587. Parent, A., L. Descarries, and A. Beaudet. Organization of ascending serotonin systems in the adult rat brain. A radioautoradiographic study after intraventricular administration of [3H]5‐hydroxytryptamine. Neuroscience 6: 115–138, 1981.
 588. Parent, A., and M. Steriade. Afferents from the periaqueductal gray, medial hypothalamus and medial thalamus to the midbrain reticular core. Brain Res. Bull. 7: 411–418, 1981.
 589. Parent, A., and M. Steriade. Midbrain tegmental projections of nucleus reticularis thalami of cat and monkey: a retrograde transport and antidromic identification study. J. Comp. Neurol. 229: 548–558, 1984.
 590. Partlow, G. D., M. Colonnier, and J. Szabo. Thalamic projections of the superior colliculus in the rhesus monkey, Macaca mulatto. A light and electron microscopic study. J. Comp. Neurol. 171: 285–318, 1977.
 591. Pasik, P., T. Pasik, and J. Hamori. Synapses between interneurons in the lateral geniculate nucleus of monkeys. Exp. Brain Res. 25: 1–13, 1976.
 592. Pasik, P., T. Pasik, J. Hamori, and J. Szentagothai. Golgi type II interneurons in the neuronal circuit of the monkey lateral geniculate nucleus. Exp. Brain Res. 17: 18–34, 1973.
 593. Passouant, P. Problèmes physiopathologiques de la narcolepsie et periodicite du “sommeil rapide” au cours du nycthemere. In: The Abnormalities of Sleep in Man, edited by H. Gastaut, E. Lugaresi, G. Berti Ceroni, and G. Coccagna. Bologna, Italy: Gaggi, 1968, p. 177–189.
 594. Pavlov, I. P. Innere Hemmung der Bedingten Reflexe und der Schlaf—ein und derselbe Prozess. Skand. Arch. Physiol. 44: 42–58, 1923.
 595. Pavlov, I. P. Conditioned Reflexes. An Investigation of the Physiological Activity of the Cerebral Cortex, translated by G. V. Anrep. New York: Dover, 1960.
 596. Payne, B. R., and N. Berman. Contralateral corticofugal projections to cat visual thalamus. Exp. Brain Res. 53: 462–466, 1984.
 597. Pearce, G. W. Some cortical projections to the midbrain reticular formation. In: Structure and Function of the Cerebral Cortex, edited by D. B. Tower and J. P. Schadé Amsterdam: Elsevier, 1960, p. 131–137.
 598. Pepeu, G., and P. Mantegazzani. Midbrain hemisection: effect on cortical acetylcholine in the cat. Science 145: 1069–1070, 1964.
 599. Peterson, B. W. Identification of reticulospinal projections that may participate in gaze control. In: Control of Gaze by Brain Stem Neurons, edited by R. Baker and A. Berthoz. Amsterdam: Elsevier, 1977, p, 143–152.
 600. Peterson, B. W., M. E. Anderson, and M. Filion. Responses of pontomedullary reticular neurons to cortical, tectal and cutaneous stimuli. Exp. Brain Res. 21: 19–44, 1974.
 601. Peterson, B. W., J. I. Franck, N. G. Pitts, and N. G. Daunton. Changes in responses of medial pontomedullary reticular neurons during repetitive cutaneous, vestibular, cortical and tectal stimulation. J. Neurophysiol. 39: 564–581, 1976.
 602. Peterson, B. W., R. A. Maunz, N. G. Pitts, and R. G. Mackel. Patterns of projection and branching of reticulospinal neurons. Exp. Brain Res. 23: 333–351, 1975.
 603. Peterson, B. W., N. G. Pitts, K. Fukushima, and R. Mackel. Reticulospinal excitation and inhibition of neck motoneurons. Exp. Brain Res. 32: 471–489, 1978.
 604. Petras, J. M. Some efferent connections of the motor and somatosensory cortex of simian primates and felid, canid and procyonid carnivores. Ann. NY Acad. Sci. 167: 469–505, 1969.
 605. Petsche, H., G. Gogolak, and P. A. Van Zwieten. Rhythmicity of septal cell discharges at various levels of reticular excitation. Electroencephalogr. Clin. Neurophysiol. 19: 25–33, 1965.
 606. Phillipson, E. A. Control of breathing during sleep. Am. Rev. Respir. Dis. 118: 909–939, 1978.
 607. Phillipson, E. A., and C. E. Sullivan. Respiratory control mechanisms during NREM and REM sleep. In: Sleep Apnea Syndromes, edited by C. Guilleminault and W. C. Dement. New York: Liss, 1978, p. 47–74.
 608. Phillipson, E. A., C. E. Sullivan, D. J. C. Read, E. Murphy, and L. F. Kozar. Ventilatory and waking responses to hypoxia in sleeping dogs. J. Appl. Physiol.: Respirat. Environ. Exercise Physiol. 44: 512–520, 1978.
 609. Phillipson, O. T. Afferent projections to the ventral tegmental area of Tsai and interfascicular nucleus: a horseradish peroxidase study in the rat. J. Comp. Neurol. 187: 117–144, 1979.
 610. Phillis, J. W. Acetylcholine release from the cerebral cortex: its role in cortical arousal. Brain Res. 7: 378–389, 1968.
 611. Phillis, J. W., and G. C. Chong. Acetylcholine release from the cerebral and cerebellar cortices: its role in cortical arousal. Nature London 207: 1253–1255, 1965.
 612. Phillis, J. W., and G. K. Kostopoulos. Activation of a noradrenergic pathway from the brain stem to rat neocortex. Gen. Pharmacol. 8: 207–211, 1977.
 613. Phillis, J. W., and D. H. York. Pharmacological studies on a cholinergic inhibition in the cerebral cortex. Brain Res. 10: 297–306, 1968.
 614. Pickel, V. M., M. Segal, and F. E. Bloom. A radioautographic study of the efferent pathways of the nucleus locus coeruleus. J. Comp. Neurol. 155: 15–42, 1974.
 615. Pivik, R. T., R. W. McCarley, and J. A. Hobson. Eye movement‐associated discharge in brain stem neurons during desynchronized sleep. Brain Res. 121: 59–76, 1977.
 616. Poitras, D., and A. Parent. Atlas of the distribution of monoamine‐containing nerve cell bodies in the brain stem of the cat. J. Comp. Neurol. 179: 699–718, 1978.
 617. Pompeiano, O. The neurophysiological mechanisms of the postural and motor events during desynchronized sleep. Proc. Assoc. Res. Nerv. Ment. Dis. 45: 351–423, 1967.
 618. Pompeiano, O. Reticular formation. In: Handbook of Sensory Physiology. Somatosensory System, edited by A. Iggo. Berlin: Springer‐Verlag, 1973, vol. II, p. 381–488.
 619. Pompeiano, O. Cholinergic activation of reticular and vestibular mechanisms controlling posture and eye movements. In: The Reticular Formation Revisited, edited by J. A. Hobson and M. A. B. Brazier. New York: Raven, 1979, p. 473–572.
 620. Pompeiano, O., and C. D. Barnes. Response of brain stem reticular neurons to muscle vibration in the decerebrate cat. J. Neurophysiol. 34: 709–724, 1971.
 621. Pompeiano, O., and K. Hoshino. Central control of posture: reciprocal discharge by two pontine neuronal groups leading to suppression of decerebrate rigidity. Brain Res. 116: 131–138, 1976.
 622. Pompeiano, O., and A. R. Morrison. Vestibular influences during sleep. I. Abolition of the rapid eye movements of desynchronized sleep following vestibular lesions. Arch. Ital. Biol. 103: 569–595, 1965.
 623. Pompeiano, O., and M. Valentinuzzi. A mathematical model for the mechanism of rapid eye movements induced by an anticholinesterase in the decerebrate cat. Arch. Ital. Biol. 114: 103–154, 1976.
 624. Porrino, L. J., and P. S. Goldman‐Rakic. Brain stem innervation of prefrontal and anterior cingulate cortex in the rhesus monkey revealed by retrograde transport of HRP. J. Comp. Neurol. 205: 63–76, 1982.
 625. Powell, T. P. S., and W. M. Cowan. The interpretation of the degenerative changes in the intralaminar nuclei of the thalamus. J. Neurol. Neurosurg. Psychiatry 30: 140–153, 1967.
 626. Powles, A. C. P., J. R. Sutton, G. W. Gray, A. L. Mansell, M. McFadden, and C. S. Houston. Sleep hypoxemia at altitude: its relationship to acute mountain sickness and ventilatory responsiveness to hypoxia and hypercapnia. In: Environmental Stress, edited by L. J. Folinsbee. New York: Academic, 1978, p. 373–381. (Proc. Int. Symp. Environ. Stress.)
 627. Puizillout, J. J., G. Gaudin‐Chazal, A. Daszuta, N. Seyfritz, and J. P. Ternaux. Release of endogenous serotonin from “encephale isolé” cats. II. Correlations with raphe neuronal activity and sleep and wakefulness. J. Physiol. Paris 75: 531–537, 1979.
 628. Puizillout, J. J., and J. P. Ternaux. Endormement vago‐aortique après section sagittale médiane du tronc cérébral et après administration de p‐chlorophenylanaline ou destruction de noyaux du raphé. Brain Res. 70: 19–42, 1974.
 629. Purpura, D. P. Interneuronal mechanisms in thalamically induced synchronizing and desynchronizing activities. In: The Interneuron, edited by M. A. B. Brazier. Berkeley: Univ. of California Press, 1969, p. 467–496.
 630. Purpura, D. P. Operations and processes in thalamic and synaptically related neural subsystems. In: The Neurosciences: Second Study Program, edited by F. O. Schmitt. New York: Rockefeller Univ. Press, 1970. p. 458–470.
 631. Purpura, D. P. Intracellular studies of synaptic organizations in the mammalian brain. In: Structure and Function of Synapses, edited by G. D. Pappas and D. P. Purpura. New York: Raven, 1972, p. 257–302.
 632. Purpura, D. P., J. G. McMurtry, and K. Maekawa. Synaptic events in ventro‐lateral thalamic neurons during suppression of recruiting responses by brain stem reticular stimulation. Brain Res. 1: 63–76, 1966.
 633. Purpura, D. P., and R. J. Shofer. Intracellular recording from thalamic neurons during reticulocortical activation. J. Neurophysiol. 26: 494–505, 1963.
 634. Purpura, D. P., R. J. Shofer, and F. S. Musgrave. Cortical intracellular potentials during augmenting and recruiting responses. II. Patterns of synaptic activities in pyramidal and nonpyramidal tract neurons. J. Neurophysiol. 27: 133–151, 1964.
 635. Rackowski, D., and I. T. Diamond. Cortical connections of the pulvinar nucleus in Galago. J. Comp. Neurol. 193: 1–40, 1980.
 636. Ramon‐Moliner, E. La différenciation morphologique des neurones. Arch. Ital. Biol. 105: 149–188, 1967.
 637. Ramon‐Moliner, E., and W. J. H. Nauta. The isodendritic core of the brain stem. J. Comp. Neurol. 126: 311–336, 1966.
 638. Ramón y Cajal, S. Histologie du système nerveux de l'homme et des vertébrés, translated by S. Azoulay. Paris: Maloine, 1909‐1911, vols. I and II.
 639. Ranck, J. B. Studies on single neurons in dorsal hippocampal formation and septum in unrestrained rats. I. Behavioral correlates and firing repertoires. Exp. Neurol. 41: 461–531, 1973.
 640. Rasminski, M., A. J. Mauro, and D. Albe‐Fessard. Projection of medial thalamic nuclei to putamen and cerebral frontal cortex in the cat. Brain Res. 61: 69–77, 1973.
 641. Reader, T. A., A. Ferron, L. Descarries, and H. H. Jasper. Modulatory role for biogenic amines in the cerebral cortex. Microiontophoretic studies. Brain Res. 160: 217–229, 1979.
 642. Rechtschaffen, A., R. A. Lovell, D. Freedman, W. E. Whitehead, and M. Aldrich. The effect of parachlorophenylalamine on sleep in the rat: some implications for the serotonin sleep hypothesis. In: Serotonin and Behavior, edited by J. Barchas and E. Usdin. New York: Academic, 1973, p. 401–418.
 643. Reed, D. J., and R. H. Kellogg. Changes in respiratory response to CO2 during natural sleep at sea level and at altitude. J. Appl. Physiol. 13: 325–330, 1958.
 644. Reed, D. Jr. and R. H. Kellogg. Effect of sleep on hypoxic stimulation of breathing at sea level and altitude. J. Appl. Physiol. 15: 1130–1134, 1960.
 645. Reed, D. J., and R. H. Kellogg. Effect of sleep on CO2 stimulation of breathing in acute and chronic hypoxia. J. Appl. Physiol. 15: 1135–1138, 1960.
 646. Reeves, A. G., and W. D. Hagamen. Behavioral and EEG asymmetry following unilateral lesions of the forebrain and midbrain in cats. Electroencephalogr. Clin. Neurophysiol. 30: 83–86, 1971.
 647. Reite, M., D. Jackson, and R. L. Cahoon. Sleep physiology at high altitude. Electroencephalogr. Clin. Neurophysiol. 38: 463–471, 1975.
 648. Reite, M., G. V. Pegram, L. M. Stephens, E. C. Bixler, and O. L. Lewis. The effect of reserpine and monoamine oxidase inhibitors on paradoxical sleep in the monkey. Psychopharrnacologia 14: 12–17, 1969.
 649. Reppert, S. M., M. J. Perlow, L. G. Ungerleider, M. Mishkin, L. Tamarkin, D. G. Orloff, H. J. Hoffman, and D. C. Klein. Effects of damage to the suprachiasmatic area of the anterior hypothalamus on the daily melatonin and Cortisol rhythms in the rhesus monkey. J. Neurosci. 1: 1414–1425, 1981.
 650. Ribak, C. E. Aspinous and sparsely‐spinous stellate neurons in the visual cortex of rats contain glutamic acid decarboxylase. J. Neurocytol. 7: 461–478, 1978.
 651. Richards, J. G., H. P. Lorez, and J. P. Tranzer. Indoleal‐kylamine nerve terminals in cerebral ventricles: identification by electron microscopy and fluorescence histochemistry. Brain Res. 57: 277–288, 1973.
 652. Richmond, B. J., and R. H. Wurtz. Constriction of visual receptive fields of inferior temporal cortex neurons during visual fixation. Invest. Ophthal. Visual Sci. 20: 148, 1981.
 653. Richter, C. P. Biological Clocks in Medicine and Psychiatry. Springfield, IL: Thomas, 1965.
 654. Rinvik, E. The corticothalamic projection from the pericruciate and coronal gyri in the cat: an experimental study with silver‐impregnation methods. Brain Res. 10: 79–119, 1968.
 655. Rinvik, E. Organization of thalamic connections from motor and somatosensory cortical areas in the cat. In: Corticothalamic Projections and Sensorimotor Activities, edited by T. Frigyesi, E. Rinvik, and M. D. Yahr. New York: Raven, 1972, p. 57–90.
 656. Rinvik, E. Demonstration of nigrothalamic connections in the cat by retrograde axonal transport of horseradish peroxidase. Brain Res. 90: 313–318, 1975.
 657. Ritvo, E. R., E. M. Ornitz, S. La Franchi, and R. D. Walter. Effects of imipramine on the sleep‐dream cycle: an EEG study in boys. Electroencephalogr. Clin. Neurophysiol. 22: 465–468, 1967.
 658. Rivner, M., and J. Sutin. Locus coeruleus modulation of the motor thalamus: inhibition in nuclei ventralis lateralis and ventralis anterior. Exp. Neurol. 73: 651–673, 1981.
 659. Robertson, R. T., and T. J. Cunningham. Organization of corticothalamic projections from parietal cortex in cat. J. Comp. Neurol. 199: 569–585, 1981.
 660. Robertson, R. T., and A. R. Feiner. Diencephalic projections from the pontine reticular formation: autoradiographic studies in the cat. Brain Res. 239: 3–16, 1982.
 661. Robertson, R. T., and S. S. Kaitz. Thalamic connections with limbic cortex. I. Thalamocortical projections. J. Comp. Neurol. 195: 501–525, 1981.
 662. Robertson, R. T., G. S. Lynch, and R. F. Thompson. Diencephalic distributions of ascending reticular systems. Brain Res. 55: 309–322, 1973.
 663. Robinson, D. A. Models of oculomotor neural organization in the control of eye movements. In: Basic Mechanisms of Ocular Motility and Their Clinical Implications, edited by G. Lennerstrand and P. Bach‐y‐Rita. Oxford, UK: Pergamon, 1975, p. 519–538. (Wenner‐Gren Ctr. Int. Symp. 1974.)
 664. Robinson, T. E., C. H. Vanderwolf, and B. A. Pappas. Are the dorsal noradrenergic bundle projections from the locus coeruleus important for neocortical or hippocampal activation? Brain Res. 138: 75–98, 1977.
 665. Roffwarg, H. P., J. N. Muzio, and W. C. Dement. Ontogenetic development of the human sleep‐dream cycle. Science 152: 604–619, 1966.
 666. Room, P., F. Postema, and J. Korf. Divergent axon collaterals of rat locus coeruleus neurons: demonstration by a fluorescent double labeling technique. Brain Res. 221: 219–230, 1981.
 667. Ropert, N., and M. Steriade. Input‐output organization of the midbrain reticular core. J. Neurophysiol. 46: 17–31, 1981.
 668. Rose, J. E., and V. B. Mountcastle. The thalamic tactile region in rabbit and cat. J. Comp. Neurol. 97: 441–489, 1952.
 669. Rossi, G. F., and A. Brodal. Corticofugal fibres to the brain stem reticular formation: an experimental study in the cat. J. Anal 90: 42–62, 1956.
 670. Rossor, M. N., C. Svendsen, S. P. Hunt, C. Q. Mountjoy, M. Roth, and L. L. Iversen. The substantia innominata in Alzheimer's disease: an histochemical and biochemical study of cholinergic marker enzymes. Neurosci. Lett. 28: 217–222, 1982.
 671. Roth, S. R., M. B. Sterman, and C. D. Clemente. Comparison of EEG correlates of reinforcement, internal inhibition and sleep. Electroencephalogr. Clin. Neurophysiol. 23: 509–520, 1967.
 672. Rougeul‐Buser, A., J. J. Bouyer, and P. Buser. From attentiveness to sleep. A topographical analysis of localized “synchronized” activities on the cortex of normal cat and monkey. Acta Neurobiol. Exp. 35: 805–819, 1975.
 673. Rougeul‐Buser, A., J. J. Bouyer, M. F. Montaron, and P. Buser. Patterns of activities in the ventrobasal thalamus and somatic cortex SI during behavioral waking immobility in the cat. Exp. Brain Res. Suppl. 7: 69–87, 1983.
 674. Roy, J. P., M. Clercq, M. Steriade, and M. Deschěnes. Electrophysiology of neurons of lateral thalamic nuclei in cat: mechanisms of long‐lasting hyperpolarizations. J. Neurophysiol. 51: 1220–1235, 1984.
 675. Roy, J. P., J. Hada, and M. Steriade. The fastigio‐ventro‐medial thalamic‐cortical projection. Soc. Neurosci. Abstr. 7: 414, 1981.
 676. Royce, G. J. Single thalamic neurons which project to both the rostral cortex and caudate nucleus studied with the fluorescent double labeling method. Exp. Neurol. 79: 773–784, 1983.
 677. Ruch‐Monachon, M. A., M. Jalfre, and W. E. Haefely. Drugs and PGO waves in the lateral geniculate body of the curarized cat. I. PGO wave activity induced by Ro 43284 and by p‐chlorophenylalanine (PCPA) as a basis for neuropharmacological studies. Arch. Int. Pharmacodyn. Ther. 219: 251–268, 1976.
 678. Ruch‐Monachon, M. A., M. Jalfre, and W. E. Haefely. Drugs and PGO waves in the lateral geniculate body of the curarized cat. II. PGO wave activity and brain 5‐hydroxytryptamine. Arch. Int. Pharmacodyn. Ther. 219: 269–286, 1976.
 679. Ruch‐Monachon, M. A., M. Jalfre, and W. E. Haefely. Drugs and PGO waves in the lateral geniculate body of the curarized cat. III. PGO wave activity and brain catecholamines. Arch. Int. Pharmacodyn. Ther. 219: 287–307, 1976.
 680. Ruch‐Monachon, M. A., M. Jalfre, and W. E. Haefely. Drugs and PGO waves in the lateral geniculate body of the curarized cat. IV. The effects of acetylcholine, GABA and benzodiapines on PGO wave activity. Arch. Int. Pharmacodyn. Ther. 219: 308–325, 1976.
 681. Ruch‐Monachon, M. A., M. Jalfre, and W. E. Haefely. Drugs and PGO waves in the lateral geniculate body of the curarized cat. V. Miscellaneous compounds. Synopsis of the role of central neurotransmitters on PGO wave activity. Arch. Int. Pharmacodyn. Ther. 219: 326–346, 1976.
 682. Rusak, B. The role of the suprachiasmatic nuclei in the generation of circadian rhythms in the golden hamster, Mesocricetus auratus. J. Comp. Physiol. 118: 145–164, 1977.
 683. Rusak, B., and G. Groos. Suprachiasmatic stimulation phase shifts rodent circadian rhythms. Science 215: 1407–1409, 1982.
 684. Rusak, B., and I. Zucker. Neural regulation of circadian rhythms. Physiol. Rev. 59: 449–526, 1979.
 685. Saito, H., K. Sakai, and M. Jouvet. Discharge patterns of the nucleus parabrachialis lateralis neurons of the cat during sleep and waking. Brain Res. 134: 59–72, 1977.
 686. Sakai, F., J. S. Meyer, I. Karacan, F. Yamaguchi, and M. Yamamoto. Narcolepsy: regional cerebral blood flow during sleep and wakefulness. Neurology 29: 61–67, 1979.
 687. Sakai, K. Some anatomical and physiological properties of pontomesencephalic tegmental neurons with special reference to the PGO waves and postural atonia during paradoxical sleep in the cat. In: The Reticular Formation Revisited, edited by J. A. Hobson and M. A. B. Brazier. New York: Raven, 1979, p. 427–447.
 688. Sakai, K., and M. Jouvet. Brain stem PGO‐on cells projecting directly to the cat dorsal lateral geniculate nucleus. Brain Res. 194: 500–505, 1980.
 689. Sakai, K., D. Salvert, M. Touret, and M. Jouvet. Afferent connections of the nucleus raphe dorsalis in the cat as visualized by the horseradish peroxidase technique. Brain Res. 137: 11–35, 1977.
 690. Sakai, K., J. P. Sastre, N. Kanamori, and M. Jouvet. State‐specific neurons in the ponto‐medullary reticular formation with special reference to the postural atonia during paradoxical sleep in the cat. In: Brain Mechanisms of Perceptual Awareness and Purposeful Behavior, edited by O. Pompeiano and C. Ajmone‐Marsan. New York: Raven, 1981, p. 405–429.
 691. Sakai, K., J. P. Sastre, D. Salvert, M. Touret, M. Tohyama, and M. Jouvet. Tegmentoreticular projections with special reference to the muscular atonia during paradoxical sleep in the cat: an HRP study. Brain Res. 176: 233–254, 1979.
 692. Sakai, K., M. Touret, D. Salvert, L. Leger, and M. Jouvet. Afferent projections to the cat locus coeruleus as visualized by the horseradish peroxidase technique. Brain Res. 119: 21–41, 1977.
 693. Sakakura, H. Spontaneous and evoked unitary activities of cat lateral geniculate neurons in sleep and wakefulness. Jpn. J. Physiol. 18: 23–42, 1968.
 694. Salmoiraghi, G. C., and B. D. Burns. Localization and patterns of discharge of respiratory neurones in brain‐stem of cat. J. Neurophysiol. 23: 1–26, 1960.
 695. Saper, C. B., L. W. Swanson, and W. M. Cowan. The efferent connections of the ventromedial nucleus of the hypothalamus of the rat. J. Comp. Neurol. 169: 409–442, 1976.
 696. Sasa, M., S. Igarashi, and S. Takaori. Influence of the locus coeruleus in interneurons in the spinal trigeminal nucleus. Brain Res. 125: 369–375, 1977.
 697. Sasa, M., and S. Takaori. Influence of the locus coeruleus on transmission in the spinal trigeminal nucleus neurons. Brain Res. 55: 203–208, 1973.
 698. Sasaki, K., H. P. Staunton, and G. Dieckmann. Characteristic features of augmenting and recruiting responses in the cerebral cortex. Exp. Neurol. 26: 369–392, 1970.
 699. Sastre, J. P., K. Sakai, and M. Jouvet. Are the gigantocellular tegmental field neurons responsible for paradoxical sleep? Brain Res. 229: 147–161, 1981.
 700. Sastre, J. P., K. Sakai, M. Tohyama, and M. Jouvet. Bilateral lesions of the dorso‐lateral pontine tegmentum. II. Effect upon muscle atonia (Abstract). Sleep Res. 7: 44, 1978.
 701. Sato, M., K. Itoh, and N. Mizuno. Distribution of thalamocaudal neurons in the cat as demonstrated by horseradish peroxidase. Exp. Brain Res. 34: 143–153, 1979.
 702. Satoh, K., M. Tohyama, T. Yamamoto, T. Sakumoto, and N. Shimizu. Noradrenaline innervation of the spinal cord studied by the horseradish peroxidase method combined with monoamine oxidase staining. Exp. Brain Res. 30: 175–186, 1977.
 703. Scheibel, A. B. The problem of selective attention: a possible structural substrate. In: Brain Mechanisms and Perceptual Awareness, edited by O. Pompeiano and C. Ajmone‐Marsan. New York: Raven, 1981, p. 319–326.
 704. Scheibel, M. E., and A. B. Scheibel. Structural substrates for integrative patterns in the brain stem reticular core. In: Reticular Formation of the Brain, edited by H. H. Jasper, L. D. Proctor, R. S. Knighton, W. C. Noshay, and R. T. Costello. Boston, MA: Little, Brown, 1958, p. 31–55.
 705. Scheibel, M. E., and A. B. Scheibel. Periodic sensory nonresponsiveness in reticular neurons. Arch. Ital. Biol. 103: 300–316, 1965.
 706. Scheibel, M. E., and A. B. Scheibel. The organization of the nucleus reticularis thalami: a Golgi study. Brain Res. 1: 43–62, 1966.
 707. Scheibel, M. E., and A. B. Scheibel. Structural organization of nonspecific thalamic nuclei and their projection toward cortex. Brain Res. 6: 60–94, 1967.
 708. Scheibel, M. E., A. B. Scheibel, and T. Davis. Some substrates for centrifugal control over thalamic cell ensembles. In: Corticothalamic Projections and Sensorimotor Activities, edited by T. Frigyesi, E. Rinvik, and M. Yahr. New York: Raven, 1972, p. 131–156.
 709. Scheibel, M., A. Scheibel, A. Mollica, and G. Moruzzi. Convergence and interaction of afferent impulses on single units of reticular formation. J. Neurophysiol. 18: 309–331, 1955.
 710. Schlag, J. D., and F. Chaillet. Thalamic mechanisms involved in cortical desynchronization and recruiting responses. Electroencephalogr. Clin. Neurophysiol. 15: 39–62, 1963.
 711. Schlag, J., I. Lehtinen, and M. Schlag‐Rey. Neuronal activity before and during eye movements in thalamic internal medullary lamina of the cat. J. Neurophysiol. 37: 982–995, 1974.
 712. Schlag, J., and J. Villablanca. Cortical incremental responses to thalamic stimulation. Brain Res. 6: 119–142, 1967.
 713. Schlag, J., and M. Waszak. Electrophysiological properties of units of the thalamic reticular complex. Exp. Neurol. 32: 79–97, 1971.
 714. Schlauch, R. Hypnopompic hallucinations and treatment with imipramine. Am. J. Psychiatry 136: 219–220, 1979.
 715. Schmielau, F. Integration of visual and nonvisual information in nucleus reticularis thalami of the cat. In: Developmental Neurobiology of Vision, edited by R. D. Freeman. New York: Plenum, 1979, p. 205–226.
 716. Schwarcz, R., T. Hökfelt, K. Fuxe, G. Jonson, M. Goldstein, and L. Ternius. Ibotenic acid‐induced neuronal degeneration: a morphobiological and neurochemical study. Exp. Brain Res. 37: 199–216, 1979.
 717. Schwartz, W. J., L. C. Davidsen, and C. B. Smith. In vivo metabolic activity of a putative circadian oscillator, the rat suprachiasmatic nucleus. J. Comp. Neurol. 189: 157–167, 1980.
 718. Schwartzkroin, P. A., and C. E. Stafstrom. Effects of EGTA on the calcium‐activated afterhyperpolarization in hippocampal CA3 pyramidal cells. Science 210: 1125–1126, 1980.
 719. Segal, M., and F. E. Bloom. The action of norepinephrine in the rat hippocampus. I. Iontophoretic studies. Brain Res. 72: 79–97, 1974.
 720. Segal, M., and F. E. Bloom. The action of norepinephrine in the rat hippocampus. II. Activation of the input pathway. Brain Res. 72: 99–114, 1974.
 721. Segal, M., and F. E. Bloom. The action of norepinephrine in the rat hippocampus. IV. The effects of locus coeruleus stimulation on evoked hippocampal unit activity. Brain Res. 107: 513–525, 1976.
 722. Segal, S., and A. J. Mandell. Behavioral activation of rats during intraventricular infusion of norepinephrine. Proc. Natl. Acad. Sci. USA 66: 289–293, 1970.
 723. Segundo, J. P., D. H. Perkel, H. Wyman, H. Hegstad, and G. Moore. Input‐output relations in computer‐simulated nerve cells. Kybernetik 4: 157–171, 1968.
 724. Segundo, J. P., T. Takenaka, and H. Encabo. Electro‐physiology of bulbar reticular neurons. J. Neurophysiol. 30: 1194–1220, 1967.
 725. Shaskan, E. G., and S. H. Snyder. Kinetics of serotonin accumulation into slices from rat brain: relationship to catecholamine uptake. J. Pharmacol. Exp. Ther. 175: 404–418, 1970.
 726. Sherrington, C. Man on His Nature. New York: Doubleday, 1955.
 727. Sheu, Y. S., J. P. Nelson, and F. E. Bloom. Discharge patterns of cat raphe neurons during sleep and waking. Brain Res. 73: 263–276, 1974.
 728. Shute, C. C. D., and P. R. Lewis. Cholinesterase containing systems of the brain of the rat. Nature London 199: 1160–1164, 1963.
 729. Shute, C. C. D., and P. R. Lewis. The ascending cholinergic reticular system: neocortical, olfactory and subcortical projections. Brain 90: 497–520, 1967.
 730. Sieck, G. C., and R. M. Harper. Pneumotaxic area neuronal discharge during sleep‐waking states in the cat. Exp. Neurol. 67: 79–102, 1980.
 731. Siegel, J. M. Behavioral functions of the reticular formation. Brain Res. Rev. 1: 69–105, 1979.
 732. Siegel, J. M., D. G. McGinty, and S. M. Breedlove. Sleep and waking activity of pontine gigantocellular field neurons. Exp. Neurol. 56: 553–573, 1977.
 733. Siegel, J. M., R. L. Wheeler, and D. J. McGinty. Activity of medullary reticular formation neurons in the unrestrained cat during waking and sleep. Brain Res. 179: 49–60, 1979.
 734. Siegfried, B., and J. Bures. Asymmetry of EEG arousal in rats with unilateral 6‐hydroxydopamine lesions of substantia nigra: quantification of neglect. Exp. Neurol. 62: 173–190, 1978.
 735. Siggins, G. R., A. P. Oliver, B. J. Hoffer, and F. E. Bloom. Cyclic adenosine monophosphate and norepinephrine effects on transmembrane properties of cerebellar Purkinje cells. Science 171: 192–194, 1971.
 736. Silberman, E. K., E. Vivaldi, J. Garfield, J. A. Hobson, and R. W. McCarley. Carbachol triggering of desynchronized sleep phenomena: enhancement via small volume infusions. Brain Res. 191: 215–224, 1980.
 737. Simonds, J. F. Enuresis. A brief survey of current thinking with respect to pathogenesis and management. Clin. Pediatr. 16: 79–82, 1977.
 738. Sims, K. B., D. L. Hoffman, S. I. Said, and E. A. Zimmerman. Vasoactive intestinal polypeptide (VIP) in mouse and rat brain: an immunocytochemical study. Brain Res. 186: 165–183, 1980.
 739. Singer, W. The effect of mesencephalic reticular stimulation on intracellular potentials of cat lateral geniculate neurons. Brain Res. 61: 35–54, 1973.
 740. Singer, W. Control of thalamic transmission by corticofugal and ascending reticular pathways in the visual system. Physiol. Rev. 57: 386–420, 1977.
 741. Singer, W. Central‐core control of visual cortex functions. In: The Neurosciences: Fourth Study Program, edited by F. O. Schmitt and F. G. Worden. Cambridge, MA: MIT Press, 1979, p. 1093–1110.
 742. Singer, W., F. Tretter, and M. Cynader. The effect of reticular stimulation on spontaneous and evoked activity in the cat visual cortex. Brain Res. 102: 71–90, 1976.
 743. Siteram, N., W. B. Mendelson, R. J. Wyatt, and J. C. Gillin. The time‐dependent induction of REM sleep and arousal by physostigmine infusion during normal human sleep. Brain Res. 122: 562–567, 1977.
 744. Siteram, N., A. M. Moore, and J. C. Gillin. Experimental acceleration and slowing of REM sleep ultradian rhythm by cholinergic agonist and antagonist. Nature London 274: 490–492, 1978.
 745. Skinner, J. E., and D. B. Lindsley. Electrophysiological and behavioral effects of blockade of the nonspecific thalamocortical system. Brain Res. 6: 95–118, 1967.
 746. Skrebitsky, V. G., and I. N. Sharonova. Reticular suppression of flash‐evoked IPSPs in visual cortex neurons. Brain Res. 111: 67–78, 1976.
 747. Snyder, F., J. A. Hobson, D. F. Morrison, and F. Goldfrank. Changes in respiration, heart rate, and systolic blood pressure in human sleep. J. Appl. Physiol. 19: 417–422, 1964.
 748. Sofroniew, M. V., and A. Weindl. Identification of parvocellular vasopressin and neurophysin neurons in the suprachiasmatic nucleus of a variety of mammals including primates. J. Comp. Neurol. 193: 659–675, 1980.
 749. Sokoloff, L., M. Reivich, C. Kennedy, M. H. Desrosiers, C. S. Patlak, K. D. Pettigrew, O. Sakurada, and M. Shinohara. The [14C]deoxyglucose method for the measurement of local cerebral glucose utilization: theory, procedure, and normal values in the conscious and anesthetized albino rat. J. Neurochem. 28: 897–916, 1977.
 750. Spencer, W. A., and J. M. Brookhart. Electrical patterns of augmenting and recruiting waves in the depths of the sensorimotor cortex of cat. J. Neurophysiol. 24: 26–49, 1961.
 751. Spencer, W. A., and J. M. Brookhart. A study of spontaneous spindle waves in sensorimotor cortex of cat. J. Neurophysiol. 24: 50–65, 1961.
 752. Sprague, J. M., M. Levitt, K. Robson, C. N. Lin, E. Stellar, and W. W. Chambers. A neuro‐anatomical and behavioral analysis of the syndromes resulting from midbrain lemniscal and reticular lesions in the cat. Arch. Ital. Biol. 101: 225–295, 1963.
 753. Stein, D., M. Jouvet, and J. F. Pujol. Effects of α‐methyl‐p‐tyrosine upon cerebral amine metabolism and sleep states in the cat. Brain Res. 72: 360–365, 1974.
 754. Steinbusch, H. W. M., D. van der Kooy, A. A. J. Verhofstad, and A. Pellegrino. Serotonergic and non‐serotonergic projections from the nucleus raphe dorsalis to the caudateputamen complex in the rat, studied by a combined immuno‐fluorescence and fluorescent retrograde axonal labeling technique. Neurosci. Lett. 19: 137–142, 1980.
 755. Stephan, F. K., K. J. Berkley, and R. L. Moss. Efferent connections of the rat suprachiasmatic nucleus. Neuroscience 6: 2625–2641, 1981.
 756. Stephan, F. K., and A. A. Nunez. Elimination of circadian rhythms in drinking, activity, sleep and temperature by isolation of the suprachiasmatic nuclei. Behav. Biol. 20: 1–16, 1977.
 757. Steriade, M. The flash‐evoked afterdischarge. Brain Res. 9: 169–212, 1968.
 758. Steriade, M. Ascending control of motor cortex responsiveness. Electroencephalogr. Clin. Neurophysiol. 26: 25–40, 1969.
 759. Steriade, M. Physiologie des voies et des centres visuels. Paris: Masson, 1969.
 760. Steriade, M. Ascending control of thalamic and cortical responsiveness. Int. Rev. Neurobiol. 12: 87–144, 1970.
 761. Steriade, M. Cortical inhibition during sleep and waking. In: Mechanisms in Transmission of Signal for Conscious Behaviour, edited by T. Desiraju. Amsterdam: Elsevier, 1976, p. 209–248.
 762. Steriade, M. Cortical long‐axoned cells and putative interneurons during the sleep‐waking cycle. Behav. Brain Sci. 1: 465–514, 1978.
 763. Steriade, M. State‐dependent changes in the activity of rostral reticular and thalamocortical elements. Neurosci. Res. Program Bull. 18: 83–91, 1980.
 764. Steriade, M. Mechanisms underlying cortical activation: neuronal organization and properties of the midbrain reticular core and intralaminar thalamic nuclei. In: Brain Mechanisms of Perceptual Awareness and Purposeful Behavior, edited by O. Pompeiano and C. Ajmone‐Marsan. New York: Raven, 1981, p. 327–377.
 765. Steriade, M. Midbrain reticular discharge related to forebrain activation processes. In: Regulatory Functions of the CNS, edited by J. Szentágothai, M. Palkovits, and J. Hámori. Oxford, UK: Pergamon, 1981, vol. 1, p. 283–292. (Adv. Physiol. Sci.)
 766. Steriade, M. Cellular mechanisms of wakefulness and slow‐wave sleep. In: Sleep Mechanisms and Functions in Humans and Animals: An Evolutionary Perspective, edited by A. Mayes. Wokingham, UK: Van Nostrand Reinhold, 1983, p. 161–216.
 767. Steriade, M. The excitatory‐inhibitory response sequence of thalamic and neocortical cells: state‐related changes and regulatory systems. In: Dynamic Aspects of Neocortical Function, edited by G. M. Edelman, W. E. Gall, and W. M. Cowan. New York: Wiley, 1984, p. 107–157.
 768. Steriade, M., V. Apostol, and G. Oakson. Control of unitary activities in cerebellothalamic pathway during wakefulness and synchronized sleep. J. Neurophysiol. 34: 389–413, 1971.
 769. Steriade, M., M. Belekhova, and V. Apostol. Reticular potentiation of cortical flash‐evoked afterdischarge. Brain Res. 11: 276–280, 1968.
 770. Steriade, M., and M. Demetrescu. Unspecific systems of inhibition and facilitation of potentials evoked by intermittent light. J. Neurophysiol. 23: 602–617, 1960.
 771. Steriade, M., and M. Deschénes. Cortical interneurons during sleep and waking in freely moving primates. Brain Res. 50: 192–199, 1973.
 772. Steriade, M., and M. Deschénes. Inhibitory processes and interneuronal apparatus in motor cortex during sleep and waking. II. Recurrent and afferent inhibition of pyramidal tract neurons. J. Neurophysiol. 37: 1093–1113, 1974.
 773. Steriade, M., and M. Deschénes. The thalamus as a neuronal oscillator. Brain Res. Rev. 8: 1–63, 1984.
 774. Steriade, M., M. Deschénes, L. Domich, and C. Mulle. Abolition of spindle oscillations in thalamic neurons disconnected from nucleus reticularis thalami. J. Neurophysiol. 54: 1473–1497, 1985.
 775. Steriade, M., M. Deschénes, and G. Oakson. Inhibitory processes and interneuronal apparatus in motor cortex during sleep and waking. I. Background firing and responsiveness of pyramidal tract neurons and interneurons. J. Neurophysiol. 37: 1065–1092, 1974.
 776. Steriade, M., M. Deschénes, P. Wyzinski, and J. Y. Halle. Input‐output organization of the motor cortex during sleep and waking. In: Basic Sleep Mechanisms, edited by O. Petre‐Quadens and J. Schlag. New York: Academic, 1974, p. 144–200.
 777. Steriade, M., A. Diallo, G. Oakson, and B. White‐Guay. Some synaptic inputs and telencephalic projections of lateralis posterior thalamic neurons. Brain Res. 131: 39–53, 1977.
 778. Steriade, M., L. Domich, and G. Oakson. Reticularis thalami neurons revisited: activity changes during shifts in states of vigilance. J. Neurosci. 6: 68–81, 1986.
 779. Steriade, M., and L. L. Glenn. Neocortical and caudate projections of intralaminar thalamic neurons and their synaptic excitation from the midbrain reticular core. J. Neurophysiol. 48: 352–371, 1982.
 780. Steriade, M., and J. A. Hobson. Neuronal activity during the sleep‐waking cycle. Prog. Neurobiol. 6: 155–376, 1976.
 781. Steriade, M., G. Iosif, and V. Apostol. Responsiveness of thalamic and cortical motor relays during arousal and various states of sleep. J. Neurophysiol. 32: 251–265, 1969.
 782. Steriade, M., A. Kitsikis, and G. Oakson. Thalamic inputs and subcortical targets of cortical neurons in areas 5 and 7 of cat. Exp. Neurol. 60: 420–442, 1978.
 783. Steriade, M., A. Kitsikis, and G. Oakson. Excitatory‐inhibitory processes in parietal association neurons during reticular activation and sleep‐waking cycle. Sleep 1: 339–355, 1979.
 784. Steriade, M., A. Kitsikis, and G. Oakson. Selectively REM‐related increased firing rates of cortical association interneurons during sleep: possible implications for learning. In: Brain Mechanisms in Memory and Learning: From the Single Neuron to Man, edited by M. A. B. Brazier. New York: Raven, 1979, p. 47–52.
 785. Steriade, M., and D. Morin. Reticular influences on primary and augmenting responses in primary somatosensory cortex. Brain Res. 205: 67–80, 1981.
 786. Steriade, M., G. Oakson, and A. Diallo. Reticular influences on lateralis posterior thalamic neurons. Brain Res. 131: 55–71, 1977.
 787. Steriade, M., G. Oakson, and A. Kitsikis. Firing rates and patterns of output and nonoutput cells in cortical areas 5 and 7 during the sleep‐waking cycle. Exp. Neurol. 60: 443–468, 1978.
 788. Steriade, M., G. Oakson, and N. Ropert. Firing rates and patterns of midbrain reticular neurons during steady and transitional states of the sleep‐waking cycle. Exp. Brain Res. 46: 37–51, 1982.
 789. Steriade, M., A. Parent, and J. Hada. Thalamic projections of nucleus reticularis thalami of cat: a study by using retrograde transport of horseradish peroxidase and fluorescent tracers. J. Comp. Neurol. 229: 531–547, 1984.
 790. Steriade, M., A. Parent, N. Ropert, and A. Kitsikis. Zona incerta and lateral hypothalamic afferents to the midbrain reticular core of cat—an HRP and electrophysiological study. Brain Res. 238: 13–28, 1982.
 791. Steriade, M., N. Ropert, A. Kitsikis, and G. Oakson. Ascending activating neuronal networks in midbrain reticular core and related rostral systems. In: The Reticular Formation Revisited, edited by J. A. Hobson and M. A. B. Brazier. New York: Raven, 1980, p. 125–167.
 792. Steriade, M., K. Sakai, and M. Jouvet. Bulbothalamic neurons related to thalamocortical activation processes during paradoxical sleep. Exp. Brain Res. 54: 463–475, 1984.
 793. Steriade, M., and P. Wyzinski. Cortically elicited activities in thalamic reticularis neurons. Brain Res. 42: 514–520, 1972.
 794. Steriade, M., P. Wyzinski, and V. Apostol. Corticofugal projections governing rhythmic activity. In: Corticothalamic Projections and Sensorimotor Activities, edited by T. Frigyesi, E. Rinvik, and M. Yahr. New York: Raven, 1972, p. 221–272.
 795. Steriade, M., P. Wyzinski, and V. Apostol. Differential synaptic reactivity of simple and complex pyramidal tract neurons at various levels of vigilance. Exp. Brain Res. 17: 87–110, 1973.
 796. Steriade, M., P. Wyzinski, M. Deschénes, and M. Guerin. Disinhibiton during waking in motor cortex neuronal chains in cat and monkey. Brain Res. 30: 211–217, 1971.
 797. Sterman, M. B., and C. D. Clemente. Forebrain inhibitory mechanisms: sleep patterns induced by basal forebrain stimulation in the behaving cat. Exp. Neurol. 6: 103–117, 1962.
 798. Stern, W. C., and P. J. Morgane. Effects of α‐methyltyrosine on REM sleep and brain amine levels in the cat. Biol. Psychiatry 6: 301–306, 1973.
 799. Stern, W. C., and P. J. Morgane. Effect of reserpine on sleep and brain biogenic amine levels in the cat. Psychopharmacologia 28: 275–286, 1973.
 800. Stone, T. W. Cholinergic mechanisms in the rat somatosensory cerebral cortex. J. Physiol. London 225: 485–499, 1972.
 801. Stone, T. W. Amino acids as neurotransmitters of corticofugal neurones in the rat: a comparison of glutamate and aspartate. Br. J. Pharmacol. 67: 545–551, 1979.
 802. Stone, T. W., D. A. Taylor, and F. E. Bloom. Cyclic AMP and cyclic GMP may mediate opposite neuronal responses in the rat cerebral cortex. Science 187: 845–847, 1975.
 803. Stool, S. E., R. D. Eavey, N. L. Stein, and W. B. Sharrar. The “chubby puffer” syndrome. Upper airway obstruction and obesity, with intermittent somnolence and cardiorespiratory embarrassment. Clin. Pediatr. 16: 43–50, 1977.
 804. Strohl, K. P., N. A. Saunders, N. T. Feldman, and M. Hallett. Obstructive sleep apnea in family members. N. Engl. J. Med. 299: 969–973, 1978.
 805. Sturmwasser, F. Neurophysiological aspects of rhythm. In: The Neurosciences: A Study Program, edited by G. C. Quarton, T. Melnechuk, and F. O. Schmitt. New York: Rockefeller Univ. Press, 1967, p. 516–528.
 806. Sugimoto, T., N. Mizuno, and K. Itoh. An autoradiographic study on the terminal distribution of cerebellothalamic fibers in the cat. Brain Res. 215: 29–47, 1981.
 807. Sweet, C. P., and J. A. Hobson. The effects of posterior hypothalamic lesions on behavioral and electrographic manifestations of sleep and waking in cat. Arch. Ital. Biol. 106: 283–293, 1968.
 808. Szabo, J. Striatonigral and nigrostriatal connections. Appl. Neurophysiol. 42: 9–12, 1979.
 809. Szentágothai, J. Basic circuitry of the neocortex. Exp. Brain Res. Suppl. 1: 282–287, 1976.
 810. Szentágothai, J., and K. Rajkovits. Der Hirnnervenanteil der Pyramidenbahn und der prämotorische Apparat motorisher Hirnnervenkerne. Arch. Psyciatr. Nervenkr. 197: 335–354, 1958.
 811. Szerb, J. C. Cortical acetylcholine release and electroencephalographic arousal. J. Physiol. London 192: 329–345, 1967.
 812. Taber, E. The cytoarchitecture of the brain stem of the cat. I. Brain stem nuclei of the cat. J. Comp. Neurol. 116: 27–70, 1961.
 813. Taber, E., A. Brodal, and F. Walberg. The raphe nuclei of the brain stem in the cat. I. Normal topography and cytoarchitecture and general discussion. J. Comp. Neurol. 114: 161–188, 1960.
 814. Taber‐Pierce, E., W. F. Foote, and J. A. Hobson. The efferent connection of the nucleus raphe dorsalis. Brain Res. 107: 137–144, 1976.
 815. Thatcher, R. W., and D. P. Purpura. Postnatal development of thalamic synaptic events underlying evoked recruiting responses and electrocortical activation. Brain Res. 60: 21–34, 1973.
 816. Tissot, R. The effects of certain drugs on the sleep cycle in man. In: Progress in Brain Research. Sleep Mechanisms, edited by K. Akert, C. Bally, and J. P. Schadé. Amsterdam: Elsevier, 1965, vol. 18, p. 175–177.
 817. Tohyama, M., K. Sakai, D. Salvert, M. Touret, and M. Jouvet. Spinal projections from the lower brain stem in the cat as demonstrated by the horseradish peroxidase technique. I. Origins of the reticulospinal tracts and their funicular trajectories. Brain Res. 173: 383–403, 1979.
 818. Tömböl, T. Collateral axonal arborization. In: Golgi Centennial Symposium: Perspectives in Neurobiology, edited by M. Santini. New York: Raven, 1975, p. 133–141.
 819. Törk, I., and S. Turner. Histochemical evidence for a cate‐cholaminergic (presumably dopaminergic) projection from the ventral mesencephalic tegmentum to visual cortex in the cat. Neurosci. Lett. 24: 215–219, 1981.
 820. Trulson, M. E., and B. L. Jacobs. Raphe unit activity in freely moving cats: correlation with level of behavioral arousal. Brain Res. 163: 135–150, 1979.
 821. Trulson, M. E., D. W. Preussler, and G. A. Howell. Activity of substantia nigra across the sleep‐waking cycle in freely moving cats. Neurosci. Lett. 26: 183–188, 1981.
 822. Udenfriend, S., P. Zaltzman‐Nirenberg, R. Gordon, and S. Spector. Evaluation of the biochemical effects produced in vivo by inhibitors of the three enzymes involved in norepinephrine biosynthesis. Mol. Pharmacol. 2: 95–105, 1966.
 823. Valverde, F. Reticular formation of the pons and medulla oblongata. J. Comp. Neurol. 116: 71–99, 1961.
 824. Valverde, F. Reticular formation of the albino rat's brain stem: cytoarchitecture and corticofugal connections. J. Comp. Neurol. 119: 25–53, 1962.
 825. Van Buren, J. M., and R. C. Borke. Variations and Connections of the Human Thalamus. New York: Springer‐Verlag, 1972, vol. 1.
 826. Vanderwolf, C. H., and T. E. Robinson. Reticulo‐cortical activity and behavior: a critique of the arousal theory and a new synthesis. Behav. Brain Sci. 4: 459–514, 1981.
 827. Vandesande, F., K. Dierickx, and J. De Mey. Identification of the vasopressin‐neurophysin producing neurons of the rat suprachiasmatic nuclei. Cell Tissue Res. 156: 377–380, 1975.
 828. Van Dongen, P. A. M. Locus coeruleus region: effects on behavior of cholinergic, noradrenergic, and opiate drugs injected intracerebrally into freely moving cats. Exp. Neurol. 67: 52–78, 1980.
 829. Vaughan, T., R. J. Wyatt, and R. Green. Changes in REM sleep of chronically anxious depressed patients given alpha‐methylparatyrosine (Abstract). Psychophysiology 9: 197, 1971.
 830. Velayos, J. L., and F. Reinoso‐Suarez. Topographic organization of the brainstem afferents to the mediodorsal thalamic nucleus. J. Comp. Neurol. 206: 17–27, 1982.
 831. Vertes, R. P. Selective firing of rat pontine gigantocellular neurons during movement and REM sleep. Brain Res. 128: 146–152, 1977.
 832. Vertes, R. P. Brain stem gigantocellular neurons: patterns of activity during behavior and sleep in the freely moving rat. J. Neurophysiol. 42: 214–228, 1979.
 833. Vibert, J. F., D. Caille, F. Bertrand, H. Gromysz, and A. Hugelin. Ascending projection from the respiratory centre to mesencephalon and diencephalon. Neurosci. Lett. 11: 29–33, 1979.
 834. Villablanca, J. The electrocorticogram in the chronic cerveau isolé cat. Electroencephalogr. Clin. Neurophysiol. 19: 576–586, 1965.
 835. Vivaldi, E., and J. A. Hobson. Cholinergic REM sleep induction by the muscarinic agonist dioxolane (Abstract). Sleep Res. 10: 104, 1981.
 836. Vivaldi, E., R. W. McCarley, and J. A. Hobson. Evocation of desynchronized sleep signs by chemical microstimulation of the pontine brain stem. In: The Reticular Formation Revisited, edited by J. A. Hobson and M. A. B. Brazier. New York: Raven, 1979, p. 513–529.
 837. Vogt, B. A., D. L. Rosene, and D. N. Pandya. Thalamic and cortical afferents differentiate anterior from posterior cingulate cortex in the monkey. Science 204: 205–207, 1979.
 838. Volterra, V. Leçons sur la théorie mathématique de la lutte pour la vie. Paris: Gauthier‐Villars, 1931.
 839. Walker, A. E. The Primate Thalamus. Chicago, IL: Univ. of Chicago Press, 1938.
 840. Wallach, M. B., W. D. Winters, A. J. Mandell, and C. E. Spooner. A correlation of EEG, reticular multiple unit activity and gross behavior following various antidepressant agents in the cat. Electroencephalogr. Clin. Neurophysiol. 27: 563–573, 1969.
 841. Wallach, M. B., W. D. Winters, A. J. Mandell, and C. E. Spooner. Effects of antidepressant drugs on wakefulness and sleep in the cat. Electroencephalogr. Clin. Neurophysiol. 27: 574–580, 1969.
 842. Walsh, J. T., and J. P. Cordeau. Responsiveness in the visual system during various phases of sleep and waking. Exp. Neurol. 11: 90–103, 1965.
 843. Walter, W. G., R. Cooper, V. J. Aldridge, W. C. McCallum, and A. L. Winter. Contingent negative variation: an electric sign of sensorimotor association and expectancy in the human brain. Nature London 203: 380–384, 1964.
 844. Waltzer, R. P., and G. F. Martin. A double labelling study demonstrating that most spinally projecting neurons of the nucleus reticularis gigantocellularis do not provide collateral innervation to the diencephalon in the rat. Soc. Neurosci. Abstr. 9: 284, 1983.
 845. Wang, R. Y., and G. K. Aghajanian. Antidromically identified serotonergic neurons in the rat midbrain raphe: evidence for collateral inhibition. Brain Res. 132: 186–193, 1977.
 846. Wang, R. Y., and G. K. Aghajanian. Physiological evidence for habenula as major link between forebrain and midbrain raphe. Science 197: 89–91, 1977.
 847. Wang, R. Y., and G. K. Aghajanian. Correlative firing patterns of serotonergic neurons in rat dorsal raphe nucleus. J. Neurosci. 2: 11–16, 1982.
 848. Wang, R. Y., D. W. Gallagher, and G. K. Aghajanian. Stimulation of pontine reticular formation suppresses firing of serotonergic neurones in the dorsal raphe. Nature London 264: 365–368, 1976.
 849. Wassef, M., A. Berod, and C. Sotelo. Dopaminergic dendrites in the pars reticulata of the rat substantia nigra and their striatal input. Combined immunocytochemical localization of tyrosine hydroxylase and anterograde degeneration. Neuroscience 6: 2125–2139, 1981.
 850. Waterhouse, B. D., and D. J. Woodward. Interaction of norepinephrine with cerebrocortical activity evoked by stimulation of somatosensory afferent pathways in the rat. Exp. Neurol. 67: 11–34, 1980.
 851. Watson, R. T., K. M. Heilman, B. D. Miller, and F. A. King. Neglect after mesencephalic reticular formation lesions. Neurology 24: 294–298, 1974.
 852. Weiss, K. R., and I. Kupfermann. Serotonergic neuronal activity and arousal of feeding in Aplysia californica. In: Aspects of Behavioral Neurobiology, edited by J. A. Ferrendelli. Washington, DC: Soc. Neurosci., 1977, vol. 3, p. 66–89. (Soc. Neurosci. Symp.)
 853. Weitzman, E. D. Sleep‐wake, neuroendocrine and body temperature circadian rhythms under entrained and non‐entrained (free‐running) conditions in man. In: Biological Rhythms and Their Central Mechanism, edited by M. Suda, O. Hayaishi, and H. Nakagawa. Amsterdam: Elsevier/North‐Holland, 1979, p. 199–227.
 854. Werner, G., and V. B. Mountcastle. The variability of central neural activity in a sensory system and its implications for the central reflection of sensory events. J. Neurophysiol. 26: 958–977, 1963.
 855. West, C. H. K., J. C. Jackson, and R. M. Benjamin. An autoradiographic study of subcortical forebrain projections from mediodorsal and adjacent midline thalamic nuclei in the rabbit. Neuroscience 4: 1977–1988, 1979.
 856. Westman, J., and D. Bowsher. Fine structure of the centromedian‐parafascicular complex in the cat. Brain Res. 30: 331–337, 1971.
 857. Wiklund, L., L. Leger, and M. Persson. Monamine cell distribution in the cat brain stem. A fluorescence histochemical study with quantification of indolaminergic and locus coeruleus cell groups. J. Comp. Neurol. 203: 613–647, 1981.
 858. Williams, J. T., R. A. North, S. A. Shefner, S. Nishi, and T. M. Egan. Membrane properties of rat locus coeruleus neurones. Neuroscience 13: 137–157, 1984.
 859. Wilson, C. J., P. M. Groves, and E. Fifkova. Monoaminergic synapses, including dendro‐dendritic synapses in the rat substantia nigra. Exp. Brain Res. 30: 161–174, 1977.
 860. Woody, C. D. Aspects of the electrophysiology of cortical processes related to the development and performance of learned motor responses. Physiologist 17: 49–69, 1974.
 861. Woody, C. D., B. E. Swartz, and E. Gruen. Effects of acetylcholine and cyclic GMP on input resistance of cortical neurons in awake cats. Brain Res. 158: 373–395, 1978.
 862. Wright, J. J., and M. D. Craggs. Changed cortical activation and lateral hypothalamic syndrome: a study in the split‐brain cat. Brain Res. 151: 632–636, 1978.
 863. Wuerthell, S. M., K. L. Lovell, M. Z. Jones, and K. E. Moore. A histological study of kainic acid‐induced lesions in the rat brain. Brain Res. 149: 489–497, 1978.
 864. Wurtz, R. H., B. J. Richmond, and W. T. Newsome. Modulation of cortical visual processing by attention, perception, and movement. In: Dynamic Aspects of Neocortical Function, edited by G. M. Edelman, W. E. Gall, and W. M. Cowan. New York: Wiley, 1984, p. 195–217.
 865. Wyatt, R. J. The serotonin‐catecholamine‐dream bicycle: a clinical study. Biol. Psychiatry 5: 33–64, 1972.
 866. Wyatt, R. J., T. N. Chase, D. J. Kupfer, J. Scott, F. Snyder, A. Sjoerdsma, and K. Engelman. Brain catecholamines and human sleep. Nature London 233: 63–65, 1971.
 867. Wyatt, R., D. J. Kupfer, J. Scott, D. S. Robinson, and F. Snyder. Longitudinal studies on the effects of monoamine oxidase inhibitors on sleep in man. Psychopharmacologia 15: 236–244, 1969.
 868. Wyzinski, P. W., R. W. McCarley, and J. A. Hobson. Discharge properties of pontine reticulospinal neurons during sleep‐waking cycle. J. Neurophysiol. 41: 821–834, 1978.
 869. Yamamoto, M., and H. Nakahama. Stochastic properties of spontaneous unit discharges in somatosensory cortex and mesencephalic reticular formation during sleep‐waking states. J. Neurophysiol. 49: 1182–1198, 1983.
 870. Yamoaka, S. Participation of limbic‐hypothalamic structures in circadian rhythm of slow wave sleep and paradoxical sleep in the rat. Brain Res. 151: 255–268, 1978.
 871. Yen, C. T., and E. G. Jones. Intracellular staining of physiologically identified neurons and axons in the somatosensory thalamus of the cat. Brain Res. 280: 148–154, 1983.
 872. Yingling, C. D., and J. E. Skinner. Selective regulation of thalamic sensory relay nuclei by nucleus reticularis thalami. Electroencephalogr. Clin. Neurophysiol. 41: 476–482, 1976.
 873. Zarcone, V. Narcolepsy. N. Engl. J. Med. 288: 1156–1166, 1973.
 874. Zatz, M., and M. J. Brownstein. Intraventricular carbachol mimics the effects of light on the circadian rhythm in the rat pineal gland. Science 203: 358–361, 1979.
 875. Zatz, M., and M. J. Brownstein. Injection of alpha‐bungarotoxin near the suprachiasmatic nucleus blocks the effects of light on nocturnal pineal enzyme activity. Brain Res. 213: 438–442, 1981.
 876. Zepelin, H., and A. Rechtschaffen. Mammalian sleep longevity and energy metabolism. Brain Behav. Evol. 10: 425–470, 1974.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

J. A. Hobson, M. Steriade. Neuronal Basis of Behavioral State Control. Compr Physiol 2011, Supplement 4: Handbook of Physiology, The Nervous System, Intrinsic Regulatory Systems of the Brain: 701-823. First published in print 1986. doi: 10.1002/cphy.cp010414