Skip to main content
Current Neuropharmacology logoLink to Current Neuropharmacology
. 2008 Sep;6(3):235–253. doi: 10.2174/157015908785777229

Functional Neuroanatomy of the Noradrenergic Locus Coeruleus: Its Roles in the Regulation of Arousal and Autonomic Function Part I: Principles of Functional Organisation

E R Samuels 1, E Szabadi 1,*
PMCID: PMC2687936  PMID: 19506723

Abstract

The locus coeruleus (LC) is the major noradrenergic nucleus of the brain, giving rise to fibres innervating extensive areas throughout the neuraxis. Recent advances in neuroscience have resulted in the unravelling of the neuronal circuits controlling a number of physiological functions in which the LC plays a central role. Two such functions are the regulation of arousal and autonomic activity, which are inseparably linked largely via the involvement of the LC. The LC is a major wakefulness-promoting nucleus, resulting from dense excitatory projections to the majority of the cerebral cortex, cholinergic neurones of the basal forebrain, cortically-projecting neurones of the thalamus, serotoninergic neurones of the dorsal raphe and cholinergic neurones of the pedunculopontine and laterodorsal tegmental nucleus, and substantial inhibitory projections to sleep-promoting GABAergic neurones of the basal forebrain and ventrolateral preoptic area. Activation of the LC thus results in the enhancement of alertness through the innervation of these varied nuclei. The importance of the LC in controlling autonomic function results from both direct projections to the spinal cord and projections to autonomic nuclei including the dorsal motor nucleus of the vagus, the nucleus ambiguus, the rostroventrolateral medulla, the Edinger-Westphal nucleus, the caudal raphe, the salivatory nuclei, the paraventricular nucleus, and the amygdala. LC activation produces an increase in sympathetic activity and a decrease in parasympathetic activity via these projections. Alterations in LC activity therefore result in complex patterns of neuronal activity throughout the brain, observed as changes in measures of arousal and autonomic function.

Key Words: Locus coeruleus, arousal, autonomic function, forebrain, diencephalons, brainstem, spinal cord, autoreceptors.

1. INTRODUCTION

The LC, a pontine nucleus located near the pontomesencephalic junction, is the largest group of noradrenergic neurones in the central nervous system [70, 174, 261]. The LC extensively projects to widespread areas of the brain and spinal cord and it was believed for many years that the outputs of this nucleus formed a diffuse and non-selective contribution to the generalised neural activation underlying themaintenance of arousal [115, 248, 268, 365, 370]. More recently, as the pathways involving the LC have been delineated, it has become clear that the projections of the LC are extremely selective [32, 105, 208, 209, 210]. In addition, the inputs to the LC are extensively varied, contributing to the complex role of this nucleus in a variety of inter-related and distinct functions.

In this review we aim to give an overview of the connections of the LC relative to two of the functions that these connections sub-serve: arousal and autonomic regulation. A number of good reviews have focused on the neuroanatomy of the LC [8, 32, 105, 243, 261, 379], but none to date have concentrated specifically on the functional pathways involving the LC in the regulation of arousal and autonomic function. It should be noted that although the LC is the preeminent noradrenergic nucleus involved in the regulation of arousal, there are some other noradrenergic nuclei, for example the A1/A5 nuclei, which also contribute to autonomic regulation. Several excellent reviews discuss these other noradrenergic nuclei in detail (for example, see 46, 65, 126, 261). In a companion review we discuss the ways in which unravelling the functional anatomy of the LC system helps with the interpretation of experimental findings involving physiological and pharmacological variables likely to act via this system (see Part II).

2. OUTPUT FUNCTIONS OF THE LOCUS COERULEUS

Noradrenergic receptors on follower cells receiving an afferent input from the LC can be generally classified as α1-, α2- or β-adrenoceptors. Activation of α1-adrenoceptors by noradrenaline generally leads to excitation of the follower cells [159] and there is some evidence that β-adrenoceptors are also excitatory [32]. In contrast, activation of α2-adrenoceptors leads to inhibition of the follower cells [159], and also of the noradrenergic neurones themselves (“autoreceptors”, see 3.2). The consequences of autoreceptors activation can be detected as changes in the firing rate of LC neurones and in the release of noradrenaline. α2 – Adrenoceptors are widely distributed in the brain [43,144], and there are regional differences in their role in modulating noradrenaline release [360].

2.1. Forebrain

2.1.1. Neocortex (Coeruleo-Cortical Pathway)

The LC extensively innervates the cerebral cortex of all hemispheric lobes [115, 161, 164, 298, 321, 370] and is the sole source of cortical noradrenaline [32, 261]. Indeed, as would be expected, there is a close correlation between LC neurone activity and noradrenaline release within the neocortex [24]. Despite the extensive nature of the projection, there is a substantial specificity within the distribution of LC fibres [32, 198, 246, 247]. It is likely that this noradrenergic input is excitatory, since α1-adrenoceptors are expressed in high concentrations throughout the cortex [75, 81, 272, 273, 281]. Interestingly, α2-adrenoceptors have also been detected in the neocortex [275, 300, 356, 358], although these tend to be fewer in number than the α1-adrenoceptors and distributed more selectively [275, 300, 356]. Surprisingly, activation of these receptors has been found to increase the activity of the neocortex [9]. It has been suggested that these α2-adrenoceptors may be present on inhibitory interneurones in the neocortex, where noradrenergic stimulation arising from the LC disinhibits the cortical neurones from inhibition by these interneurones and thus leads to an increase in cortical activity [9].

The noradrenergic projection to the neocortex is likely to contribute to the generally recognised role of the LC as a major wakefulness-promoting nucleus (for example, see 256, 257). A number of pieces of converging evidence support this role. Firstly, the activity of the LC closely correlates with level of arousal [14, 15, 16, 103, 104, 287, 291]. Secondly, increases in LC activity have been found to increase EEG signs of cortical arousal [25] whilst LC inactivation results in a reduction in cortical EEG activity [30, 73]. Indeed, increasing the noradrenergic activation of α1-adrenoceptors in the prefrontal cortex in rats has been found to increase cognitive performance, and this was interpreted as resulting from an increase in arousal [191]. Finally, electrical stimulation of the LC in a human subject resulted in a reduction in the quantity of both slow wave and rapid eye movement sleep [165], again supporting a wakefulness-promoting role for LC activation.

2.1.2. Basal Forebrain

The BF, comprising the medial septal area, medial preoptic area and substantia innominata, contains both cholinergic and GABAergic cortically-projecting neurones involved in modulating the sleep-wakefulness state [122, 159, 242]. All three areas of the BF receive noradrenergic inputs from the LC [96, 163, 406]. The cholinergic BF neurones are most active during wakefulness, corresponding to an increase in cortical activation, whilst the GABAergic BF neurones are most active during slow-wave sleep and show reduced activity when cortical activation is high [159, 160, 221, 222]. The cholinergic wakefulness-promoting neurones of the BF are activated by noradrenaline released from the terminals of neurones projecting from the LC [96, 106, 159, 160] and this activation is likely to result from stimulation of α1- and β-adrenoceptors identified in the medial preoptic and septal areas [26, 27, 31, 106]. Indeed, pharmacological activation of the of α1- and β-adrenoceptors in the medial preoptic and septal areas results in an additive increase in arousal above the level achieved by the stimulation of either receptor type alone [27]. In contrast, the GABAergic sleep-promoting neurones of the BF, situated in the medial preoptic area, are inhibited by the stimulation of α2-adrenoceptors [223, 242, 356]. Thus, the noradrenergic projection from the LC inhibits these sleep-promoting neurones to promote wakefulness, and these neurones become disinhibited following LC quiescence at the onset of sleep [159, 160, 223]. The overall effect of LC innervation to the cholinergic and GABAergic neurones of the BF is therefore the promotion of arousal. In agreement with this, infusions of noradrenaline into the BF in the rat increase signs of wakefulness [28, 29, 50, 159].

2.1.3. Limbic System

The amygdala is principally responsible for fear and anxiety responses to threatening environmental stimuli, including the increase in the activity of the sympathetic nervous system to threat [71, 174, 192]. The LC densely innervates the amygdala [164, 243] and in particular projects to the central and basal nuclei [98, 163, 261, 354]. The amygdala primarily expresses α1-adrenoceptors [75, 81, 272, 273, 281], although α2-adrenoceptors have also been detected within this area [117, 275, 300, 356]. These α2-adrenoceptors may be located on particular subsets of neurones involved in the autonomic response to stressful stimuli (for example, see 117). The overall noradrenergic influence on the amygdala, however, is likely to be largely excitatory.

Activation of the LC by electrical stimulation or drug administration (for example yohimbine) results in observations of increased anxiety [57, 118, 234, 252, 293, 294] [368], likely as a result of the potentiation of this excitatory pathway from the LC to the amygdala. In addition to a role in anxiety, the LC projection to the amygdala may also play a role in forming and retrieving emotional memories [59, 340]. Interestingly, level of arousal, highly correlated to LC activity (see 2.1.1), determines the likelihood of a memory being encoded and subsequently retrieved. Indeed, both α1- and β-adrenoceptors in the basolateral amygdala have been implicated in memory storage [100, 101].

The LC also innervates the hippocampus, providing the sole source of noradrenaline to the hippocampal neurones [110, 174, 204, 208, 210, 261, 277, 369]. Both α1-adrenoceptors [271, 281], particularly α1D-adrenoceptors [75], and α2-adrenoceptors [275, 300] have been detected in the hippocampal formation. This limbic structure is centrally involved in the formation of declarative memories and the LC projection to the hippocampus may thus contribute to memory formation. In support of this possible function, lesions of the LC impair olfactory learning in experimental animals [343]. The LC has also been implicated in memory retrieval [80, 311], although this is likely to be mediated through areas outside of the hippocampus, possibly involving the amygdala [174] (see above).

2.2. Diencephalon

2.2.1. Thalamus

The LC sends a large output to the thalamus, especially to the dorsally located nuclei [147, 164, 261]. The principle adrenoceptor identified in the thalamus appears to be the excitatory α1-adrenoceptor [75, 281, 409]. Some studies have also reported the presence of α2-adrenoceptors in the thalamus [356, 358, 409], although others have not [275, 300]. This output to the thalamus may be related to the wakefulness-promoting role of the LC (see 2.1.1), since thalamic neurones project extensively to the cortex [158, 231]. Indeed, noradrenergic input to the thalamic cells promotes a single spike firing mode of activity in the thalamus that has been related to increased cortical activity and responsiveness during waking [231, 233]. In addition to a role in wakefulness, sparse projections from the LC to the ventral posterolateral nucleus of the thalamus may be involved in modulating the sensation of pain [389].

2.2.2. Hypothalamus

2.2.2.1. Ventrolateral Preoptic Area

In addition to the cortical and thalamic projections described above, the LC contributes to the maintenance of arousal via an inhibitory output to the GABAergic neurones of the VLPO of the hypothalamus [62, 242, 270], an area associated with the regulation of sleep-wakefulness state (see Figs. (1) and (4) in Part II). Neurones of the VLPO exhibit a high activity state during SWS and REM sleep, whilst being virtually silent during periods of waking [354]. This is in contrast to the activity of the LC, where neurones are active during wakefulness, quiet during SWS and quiescent during REM sleep [14, 16, 104, 287]. Noradrenaline, released from the LC during wakefulness, inhibits the majority of VLPO neurones through the stimulation of α2-adrenoceptors [114, 228, 241, 270]. When active the VLPO inhibits multiple areas involved with promoting wakefulness, primarily the TMN of the hypothalamus, through GABAergic projections [324]. The TMN sends wakefulness-promoting histaminergic projections to the cerebral cortex [181]; inhibition of the VLPO by the LC thus disinhibits this wakefulness-promoting projection from the TMN. Through this projection to the VLPO and the excitatory projections to the cortex and thalamus (see 2.1.1 and 2.2.1) the LC plays an important role in maintaining arousal [256, 257].

2.2.2.2. Paraventricular Nucleus

A second hypothalamic area to receive a significant projection from the LC is the PVN, a major premotor autonomic nucleus associated with both the sympathetic and parasympathetic nervous systems [163, 164, 261, 350, 351]. Autonomic projections from the PVN terminate on sympathetic preganglionic neurones in the IML of the spinal cord [35, 310, 317, 350, 377] and on parasympathetic preganglionic nuclei in the brainstem [35, 44, 137, 317, 391, 409]. Expression of α1-adrenoceptors has been detected in the PVN [3, 75, 273, 281, 309, 342] and the activation of these receptors has been linked to the autonomic response to stress [309, 342]. In addition, α2-adrenoceptors have been detected on inhibitory GABAergic neurones synapsing with these spinally-projecting PVN neurones, providing a further means of activating the autonomic nervous system through noradrenaline release from the LC [200]. Through this projection, therefore, the LC influences functions of the autonomic nervous system relating to behavioural arousal, for example cardiovascular regulation (suppression of the baroreceptor reflex: 145, 326).

The LC is also involved in neuroendocrine function by projections to the neuroendocrine cells of the PVN and tuberoinfundibular nucleus of the hypothalamus (see 2.2.2.4). These neuroendocrine cells are responsible for the secretion of trope hormones, for example thyrotropin-releasing hormone and somatostatin, modulating pituitary activity (see 305).

2.2.2.3. Lateral Hypothalamus/Perifornical Area

The orexin (also called hypocretin) neurones of the LH/PF receive a noradrenergic input from the LC [18, 202, 398]. In contrast to the excitatory influence of orexin on the LC (see 3.1.3.3, below), this noradrenergic input inhibits LH/PF firing [202, 398]. This suggests that there is a negative feedback circuit from the LC to the LH/PF, likely to be involved in preventing excessive activity in the arousal pathway during periods of wakefulness. Some excitatory α1-adrenoceptors have also been identified in the LH/PF [3, 281, 342], where activation of these receptors has been linked to behavioural activation and exploration [342]. However, many reports on α1-adrenoceptor localisation do not detect significant numbers in the LH/PF [75, 81, 272]. Interestingly, β-adrenoceptors in the LH/PF appear to be related to noradrenergic suppression of feeding [51, 184].

2.2.2.4. Tuberoinfundibular Area

The LC projects to the arcuate nucleus of the tuberoinfundibular area, the neurones of which are involved in neuroendocrine regulation [123, 167]. Both excitatory α1- and β-adrenoceptors [3, 167] and inhibitory α2-adrenoceptors [404] have been detected in the arcuate nucleus. The LC may contribute to the regulation of neuroendocrine function via these adrenoceptors along with the projection to the PVN (see 2.2.2.2). For example, the arcuate nucleus neurones controlling the secretion of GHRH, which increases GH release from the pituitary [249], are modulated by α2-adrenoceptor stimulation [64, 404]. Similarly, the arcuate nucleus is involved in the regulation of prolactin secretion via the release of dopamine onto the lactotropes in the pituitary [23, 295], which inhibits prolactin secretion. Noradrenergic stimulation of α1-adrenoceptors activates these dopaminergic arcuate nucleus neurones and thus contributes to the regulation of prolactin secretion [141]. Indeed, administration of modafinil, a wakefulness-promoting drug believed to enhance noradrenergic LC activity [139] and therefore noradrenaline release onto the dopaminergic neurones, was found to reduce prolactin secretion in healthy male volunteers [305].

2.3. Brainstem

2.3.1. Parasympathetic Preganglionic Nuclei

In general, the LC inhibits parasympathetic preganglionic nuclei (Edinger-Westphal nucleus, salivatory nuclei, and vagal nuclei) via the activation of inhibitory α2-adrenoceptors on the neuronal membrane of the preganglionic follower cells (see Figs. (2) and (4) in Part II).

2.3.1.1. Edinger-Westphal Nucleus (Coeruleo-Pupillomotor Pathway)

The EWN, the parasympathetic preganglionic nucleus responsible for pupil constriction, receives an ascending input from the LC [42, 207], which is likely to exert an inhibitory influence via α2-adrenoceptors [133, 134, 352](see Fig. (2) in Part II). In the pathway controlling pupil constriction, light stimulation is detected at the retina and transmitted via the pretectal nucleus to the EWN [205]. The EWN, located in the oculomotor complex of the midbrain, innervates the ciliary ganglion supplying the sphincter pupillae muscle: activation of this muscle results in constriction of the pupil and a reduction in the level of luminance entering the eye. Through this pathway the EWN mediates both the constriction of the pupil in situations of high background luminance and the reflex constriction of the pupil to a sudden increase in luminance level (light reflex response) [175, 181].

The functional significance of the projection from the LC to the EWN is highlighted by two different observations. Firstly, the activation of the LC attenuates the light reflex response (see 2.2.2, Part II). Secondly, there is a species difference in the response of the pupil to α2-adrenoceptor agonists, likely to result from a differential activation of α2-adrenoceptors within the LC and EWN (see 3.1.1.7, Part II).

2.3.1.2. Salivatory Nuclei (Coeruleo-Salivatory pathway)

The salivatory nuclei are located in the reticular formation, with cells situated ventrolaterally designated as the inferior salivatory nucleus and cells situated dorsolaterally designated as the superior salivatory nucleus [259]. The inferior and superior salivatory nuclei are so divided according to their projections to the periphery: the inferior salivatory nucleus projects through the glossopharyngeal nerve whilst the superior salivatory nucleus projects through the facial nerve. The inferior salivatory nucleus is responsible for the parasympathetic innervation of the otic ganglion, from which postganglionic fibres innervate the parotid and lingual (von Ebner) salivatory glands involved in salivation and taste perception respectively [41, 111, 175]. The superior salivatory gland is responsible for parasympathetic innervation of the submandibular ganglion, from which postganglionic fibres innervate the submandibular and sublingual salivatory glands [61, 151, 319, 405] and the pterygopalatine ganglion innervating the lacrimal gland involved in tear secretion [258, 365]. The neurones of the superior salivatory nucleus may be divided into two categories: type I neurones are responsible for salivation whilst type II neurones are involved in vasodilatation [229].

There is limited information regarding the innervation of the salivatory nuclei, but that which is available suggests a possible role for the LC in modulating these nuclei, and thus contributing to the autonomic control of salivation. The neurones of the superior salivatory nuclei receive synaptic inputs from monoaminergic cell groups [258] and within this there may be a noradrenergic contribution originating in the LC [151, 338]. Inhibitory α2-adrenoceptors are located on the preganglionic parasympathetic neurones of the salivatory nuclei [214, 353] and it has been suggested that salivation is tonically inhibited via the activation of these receptors [355]. Indeed, there is pharmacological evidence consistent with the existence of inhibitory α2-adrenoceptors on salivatory neurones (see 3.1.1.4, Part II).

2.3.1.3. Parasympathetic Vagal Nuclei (Coeruleo-Vagal Pathway)

The parasympathetic vagal nuclei include the DMV and the nucleus ambiguus. The DMV, the largest preganglionic parasympathetic nucleus in the brainstem, has efferents contributing to the control of smooth muscle in the thoracic and abdominal viscera [174]. Along with the nucleus ambiguus (see below), the DMV mediates the parasympathetic control of cardiovascular function [264]. The neurones of the DMV fire in synchrony with the cardiac cycle to reduce heart rate and blood pressure: excitatory cyclical input from the baroreceptors, mediated through the nucleus of the solitary tract, increases the firing of neurones in the DMV [72]. Indeed, microinjections of glutamate into the DMV were found to reduce both heart rate and blood pressure [58]. The LC projects to the DMV [261, 359, 388] and α2-adrenoceptors have been detected on DMV neurones [275, 299, 300, 370]. It has been shown that noradrenaline inhibits the activity of DMV neurones via the activation of these α2-adrenoceptors [112, 226]. Furthermore, microinjection of glutamate into the LC was found to increase HR and BP [58], consistent with an inhibitory influence of the LC on the DMV. This observation is also consistent with the combined sympathomimetic and parasympatholytic effect of LC activation on cardiovascular function.

The nucleus ambiguus innervates the cardiac ganglia to contribute to the control of heart rate [174, 218] and is critical for the heart rate response to baroreceptor stimulation [60]. Chemical stimulation of the neurones of the nucleus ambiguus results in a reduction in heart rate [225] and blood pressure [58]. The nucleus ambiguus also reduces heart rate through an inhibitory projection to the rostroventrolateral medulla [236], the major pre-sympathetic nucleus involved in cardiovascular regulation (see 2.3.2.1). The LC projects to the nucleus ambiguus [164, 388] and α1-adrenoceptors have been detected within this nucleus [75]. The presence of α1-adrenoceptors in this nucleus appears to be paradoxical since α1-adrenoceptors usually mediate excitation and the LC generally exerts an inhibitory effect on preganglionic cholinergic neurones. However, the cellular localisation of these α1-adrenoceptors has not been identified and it is possible that they are located on inhibitory interneurones rather than on the preganglionic cholinergic output neurones themselves, as would be consistent with the general pattern of the modulation of autonomic activity by the LC.

2.3.2. Premotor Sympathetic Nuclei

2.3.2.1. Rostroventrolateral Medulla (Coeruleo-Vasomotor Pathway)

In addition to the modulation of parasympathetic nuclei (see 2.3.1.3), the LC contributes to the regulation of cardiovascular function via inhibitory projections to the RVLM [376] (see Figs. (2) and (4) in Part II). The RVLM tonically projects to sympathetic preganglionic neurones in the IML of the spinal cord [19, 53, 72, 344, 346, 407], where glutamate release excites the neurones and promotes vasoconstriction. Some of the spinal-projecting RVLM neurones have an intrinsic pacemaker activity that contributes to the maintenance of normal blood pressure and heart rate [72, 125, 344]. The RVLM neurones are also involved in the response of cardiovascular activity to changing environmental demands, mediated via the baroreflex. Baroreceptor activation triggered by an increase in blood pressure enhances nucleus of the solitary tract neurone activity, which activates the GABAergic caudal ventrolateral medulla neurones that inhibit RVLM activity [48, 72, 344, 345, 346]. The activation of this pathway mediates a reduction in cardiovascular activity. In this way, RVLM neuronal activity is synchronised to the cardiac-related rhythm in sympathetic activity [19], with reduced firing rate following baroreceptor activation. In contrast, hypotension increases RVLM activity [119]. In general, chemical or electrical stimulation of the RVLM produces an increase in blood pressure and heart rate, while chemical lesions of the RVLM reduce blood pressure [48, 72, 125, 168, 344, 392]. The inhibition of the RVLM by the LC is likely to result from the stimulation of α2-adrenoceptors located within the RVLM [132, 173, 319, 358]. For the functional significance of these receptors see section 3.1.1.3, Part II.

The combined influences of the LC on the sympathetic output to the cardiovascular system (coeruleo-vasomotor and coeruleo-spinal pathways) result in a moderate increase in blood pressure and heart rate when the LC is activated [82, 124, 203, 331, 341, 385], indicating the predominance of the direct innervation of the spinal cord (see 2.5.3). The inhibition of the DMV and nucleus ambiguus (coeruleo-vagal pathway) may also contribute to this effect (see 2.3.1.3). Hypertension has been found to increase GABA release in the LC [332], leading to a reduction in LC neurone activity [248] and a decrease in noradrenaline release [332]. In contrast, hypotension has been found to decrease GABA release in the LC [332], leading to an increase in LC neurone activity [10, 362, 372] and an increase in noradrenaline release [332]. It thus appears that the excitation of the spinal cord and possibly the inhibition of the DMV and nucleus ambiguus by the LC predominate over the inhibition of the RVLM.

2.3.2.2. Caudal Raphe Nuclei

The nuclei of the CR (raphe magnus, obscurus, and pallidus) are innervated by noradrenergic projections from the LC [136, 164], which are likely to have primarily excitatory effects via α1-adrenoceptor activation [81, 271]. However, α2-adrenoceptors have also been identified within the CR [127, 300, 370], suggesting that a subset of CR neurones may be inhibited by the LC projection. The CR is involved in modulating sympathetic function via serotonergic outputs to the IML of the spinal cord [7, 150, 206]. Both the raphe pallidus and the raphe magnus have been implicated as excitatory premotor neurones in the regulation of body temperature, innervating sympathetic preganglionic neurones in the IML [255]. This is supported by studies finding that suppression of the activity of the raphe pallidus with microinjection of muscimol, a GABA receptor agonist, results in hypothermia [408] and that cold exposure increases neurone activity within the CR, measurable both as an increase in electrical unit activity [292] and as an increase in Fos expression [36, 245, 254]. Therefore, the excitation of these CR neurones by the projection from the LC would be expected to increase body temperature via the activation of the sympathetic nervous system.

Another major action of the serotonergic projection to the spinal cord is the suppression of nociception [66, 113, 239], achieved through the release of serotonin in the dorsal spinal cord [135, 162, 188, 297]. Interestingly, α1- and α2-adrenoceptors have been found to be co-localised on raphe magnus neurones involved in the inhibition of nociception [34]. These receptors have been implicated in the mechanism underlying opioid-induced analgesia, since blocking the α1-adrenoceptors or stimulating the α2-adrenoceptors attenuated the analgesia produced through local opioid administration [34]. Thus, the innervation by the LC is likely to be acting at excitatory α1-adrenoceptors to increase CR neurone activity and thus contribute to the suppression of nociception during opioid analgesia.

The projection from the LC to the CR is therefore involved in modulating both the premotor sympathetic neurones involved in thermoregulation and the neurones involved in nociception.

2.3.3. Dorsal Raphe Nucleus

In addition to the projection to the neurones of the CR described above (see 2.3.2.2), the LC also innervates the serotonergic neurones of the DR nucleus [176, 197, 224, 261, 304]. The DR is involved in the regulation of the sleep-wakefulness state and DR serotonergic neurones have been found to fire extensively during wakefulness whilst showing quiescence during periods of sleep [235, 366]. The input from the LC to the DR appears to be excitatory via activation of α1-adrenoceptors [43, 75, 76, 271, 272, 281, 284, 342, 402] and it may be involved in the maintenance of increased DR neurone firing during wakefulness. Indeed, noradrenaline perfusion directly into the DR results in cortical desynchronisation [172].

2.3.4. Pedunculopontine and Laterodorsal Tegmental Nuclei

The cholinergic neurones of the PPT and LDT are associated with the regulation of the sleep-wakefulness state. The neurones of the PPT and LDT are active during both wakefulness and REM sleep [89, 160, 171]. These neurones project to the thalamus, where they excite neurones involved in promoting cortical desynchrony [160, 169]. Two populations of neurones have been proposed within these nuclei: one set of cholinergic neurones are active during waking and excited by noradrenaline from the LC acting at α1-adrenoceptors, and one set of cholinergic neurones that are active during REM sleep and inhibited by noradrenaline from the LC acting at α2-adrenoceptors [139, 140]. This second population of neurones may be largely responsible for the initiation of REM sleep [303]; during wakefulness REM sleep is inhibited by the activation of α2-adrenoceptors on these neurones [20].

2.3.5. Motor Nuclei

The LC projects to motoneurones in the brainstem and the spinal cord, facilitating motoneurone activity via the stimulation of α1-adrenoceptors.

2.3.5.1. Facial Nucleus

The LC densely projects to the motoneurones of the facial nucleus [164, 231], a group of motoneurones located at the pontomedullary junction, and this projection appears to be excitatory since extracellular microiontophoretic application of noradrenaline increases the activity of these motoneurones [289, 378, 390]. Indeed, excitatory α1-adrenoceptors have been detected within the facial nucleus [75, 81], supporting this facilitatory action.

There is evidence from studies using the acoustic startle paradigm that the facilitatory projection from the LC to the facial nucleus may be tonically active. The acoustic startle paradigm involves the presentation of a sudden intense auditory stimulus to produce a startle response involving the rapid involuntary contraction of facial and skeletal musculature. The conventional measure of this response is the EMG recording of the orbicularis oculi muscle, involved in the eye blink response, which is innervated by the motoneurones of the facial nucleus. It has been found that the administration of a sedative drug such as clonidine, known to reduce LC activity, results in a reduction in the amplitude of the acoustic startle response [1, 2, 187, 307], whilst the administration of the α2-adrenoceptor antagonist yohimbine [148, 333, 357], known to increase LC activity, enhances the amplitude of the response [244].

2.3.5.2. Hypoglossal Nucleus

There is limited information regarding the afferents of the hypoglossal nucleus, but there is some evidence of a noradrenergic influence on hypoglossal motoneurones. Retrograde and anterograde transport techniques have identified a projection to the hypoglossal nucleus from the subcoeruleus nucleus [6], which may include neurones of the LC. Indeed, the descending projection of the LC passes ventrolaterally to the hypoglossal nucleus [231] and thus it is possible that fibres from this pathway innervate the nucleus. Additionally, there are α1-adrenoceptors, but not α2-adrenoceptors, on these neurones [274, 343] suggesting an excitatory influence of noradrenaline on the nucleus. Application of noradrenaline to the hypoglossal nucleus results in motoneurone activation and the application of an α1-adrenoceptor agonist (phenylephrine) mimicked this effect [274]. In contrast, the application of an α1-adrenoceptor antagonist (prazosin) prevented the noradrenaline-induced increase in activity [274].

2.3.5.3. Trigeminal Motor Nucleus

Although the majority of noradrenergic projections to the trigeminal motoneurones arise from the A5/A7 nuclei, the LC also projects to this nucleus, albeit sparsely [217]. In general, noradrenaline has a facilitatory influence on these motoneurones [323]. Bilateral locus coeruleus lesions do not significantly reduce noradrenaline content in the trigeminal motor nucleus, suggesting that this input from the LC is of minor physiological significance [197].

2.3.5.4. Oculomotor Nuclear Complex

The motor neurones situated in the nuclei of the third (oculomotor), fourth (trochlear), and sixth (abducens) cranial nerves form the oculomotor nuclear complex responsible for innervating the external muscles of the eyes controlling the movements of the eye. A small number of cells in the LC have been found to project to the oculomotor nuclear complex [52] and high levels of α1-adrenoceptors have been identified within this area [75, 81, 281], indicating an excitatory noradrenergic input to these neurones.

2.3.6. Sensory Nuclei

2.3.6.1. Trigeminal Sensory Nucleus

In contrast to the sparse innervation of the trigeminal motor nucleus, the LC densely innervates the neurones of the trigeminal sensory nucleus [66, 197, 261, 321] and this pathway is likely to be involved in the antinociceptive function of the LC [47, 66, 367]. Indeed, electrical stimulation of the LC has been found to inhibit the neurones of the sensory trigeminal nucleus involved in pain perception [230, 312, 314, 367]. In addition, inhibitory α2-adrenoceptors have been detected within the trigeminal nucleus [410]. It has been proposed that β-adrenoceptors may also be involved in the inhibitory influence of noradrenaline on the trigeminal sensory neurones [313]. Recently it has been shown that there is an intricate synergistic interaction between the antinociceptive effects of the noradrenergic and serotonergic inputs to the sensory trigeminal nucleus [66].

2.3.6.2. Cochlear Nucleus

The LC projects diffusely to the cochlear nuclei and noradrenaline levels are detectable here in moderate concentrations [180, 185, 188, 320, 341]. This input is suggested to be excitatory, where noradrenaline enhances both spontaneous and tone-evoked cochlear nucleus neurone activity, and it is likely that this effect is mediated via α1-adrenoceptors [84]. The LC may, therefore, be involved in sensory auditory processing.

Noradrenergic modulation of cochlear nucleus activity may underlie the mechanism by which clonidine reduces the amplitude of auditory evoked potentials (N1/P2) recorded in the EEG during the acoustic startle paradigm [307] since clonidine is known to act by reducing LC activity. It should be noted, however, that this effect of clonidine on the auditory evoked potentials has not been observed in every study [2].

2.4. Cerebellum (Coeruleo-Cerebellar Pathway)

The cerebellum is responsible for the planning, coordination, and learning of movements, particularly relating to the timing, force and extent of muscle contractions, and may also be involved in cognition and emotion [174]. The LC projects to areas throughout the cerebellum [208, 210, 261, 268, 369] and in particular to the cerebellar cortex [301, 320]. A moderate number of α1-adrenoceptors has been observed in the cerebellum [75, 271, 273, 318, 342], indicating an excitatory role for the LC in facilitating one or more of the functions of the cerebellum. Indeed, depletion of noradrenaline from the cerebellum has been found to result in impaired motor performance [386].

2.5. Spinal Cord (Coeruleo-Spinal Pathway)

The contribution of the LC to autonomic nervous system control involves a direct output to sympathetic and parasympathetic preganglionic neurones of the IML of the spinal cord in addition to the projections innervating other autonomic nuclei, for example the EWN, and premotor autonomic nuclei, for example the PVN, CR, and RVLM, described above (see 2.3.1.1, 2.2.2.2 and 2.3.2.2; see also Figs. (2) and (4) in Part II). The LC also contributes to sensory and motor functions through projections to the dorsal and ventral horns of the spinal cord, respectively. The LC projects to all three areas (dorsal and ventral horns and IML) to differing extents [131, 164, 196, 38] and there is some evidence that this distribution may differ between rat strains [283, 334].

2.5.1. Dorsal Horn

Neurones in the dorsal horn of the spinal cord are sensory neurones associated with the detection of pain, temperature, touch and position/movement (proprioception) [174]. The LC most densely innervates the cells in this compartment of the spinal cord [63, 107, 108, 265, 283, 387, 388], signifying that the LC can influence the processing of sensory information. This influence is likely to be inhibitory, achieved via the activation of pre-synaptic α2-adrenoceptors on the terminals of excitatory peptidergic spinal interneurones within the dorsal horn [267]. In support of this suggestion α2-adrenoceptors have been identified within this region [189, 275, 325, 336]. The projection from the LC may be particularly important in the processing of noxious stimuli (nociception). Indeed, it has been suggested that the spinal projection from the LC has analgesic properties [331] and α2-adrenoceptor agonists acting in the dorsal horn of the spinal cord are known to produce analgesia [88]. Interestingly, α2-adrenoceptor antagonists increase the response of dorsal horn neurones to inflammation induced by formalin injection [121]. Additionally, β-adrenoceptors have been detected in the dorsal horn, and may also be involved with nociception [241]. The importance of noradrenaline in the modulation of nociception has been highlighted by the observation of hyperalgesia in the absence of noradrenaline in mice lacking the gene coding for the noradrenaline-synthesising enzyme dopamine-β-hydroxylase [152, 335].

2.5.2. Ventral Horn

The LC sends projections to the neurones of the ventral horn [164, 265, 387], innervating the skeletal musculature, and this pathway may contribute to muscle contraction and tone. Indeed, excitatory α1-adrenoceptors are present on the motoneurones of the ventral horn [75, 81, 337], supporting a facilitatory influence of the LC on muscle tone. Both α2-adrenoceptors [336] and β-adrenoceptors [241] have also been detected in this region, although the role of these receptors is not clear at present. The loss of LC activity during attacks of cataplexy in narcolepsy [396] may explain the sudden loss of muscle tone characteristic of these attacks [362].

2.5.3. Intermediolateral Cell Column

Neurones of the IML form separate sympathetic preganglionic nuclei that project to ganglia innervating specific target organs. Through these nuclei the sympathetic nervous system can be selectively activated, as opposed to generalised sympathetic activation [11]. The LC has been found to project to the cells of the IML [164, 265, 387] where noradrenaline excites the majority of sympathetic preganglionic neurones [199], possibly via the activation of α1-adrenoceptors detected in this region [281]. The densest projections from the LC to the IML, however, end in the sacral spinal segments where parasympathetic inhibitory interneurones are located [164, 315, 388, 403] and are involved in functions such as micturition [315]. The projection from the LC to the IML, results in excitation of these inhibitory interneurones via the stimulation of α1-adrenoceptors [315, 403], which, in turn, leads to a reduction in parasympathetic outflow. There may also be a direct inhibitory influence on the parasympathetic neurones of the IML, since α2-adrenoceptors are present within the lumbosacral parasympathetic segments [336, 370, 397].

3. MODULATION OF LOCUS COERULEUS ACTIVITY

As has become clear, the projections from the LC to the many widespread areas of the neuraxis are complex and extensive. To increase the complexity surrounding this nucleus, the LC also receives multiple varied inputs, which all influence LC firing to differing extents. In the majority of instances the neurotransmitter involved in these inputs to the LC is known; in some however it remains uncertain.

3.1. Modulation by Heteroreceptors via Afferent Inputs

A detailed review of the neuroanatomical techniques used to identify afferent inputs to the LC is beyond the scope of this paper, although some of the methods used are covered in Part II.

3.1.1. Neocortex

Although in general the parietal, temporal, infralimbic, insular and frontal cortices provide only a limited input to the LC [12, 54, 215], there is a strong reciprocal connection between the LC and the prefrontal cortex [153, 154, 332], a cortical area involved in executive functioning. Indeed, the LC has been found to contribute to the regulation of functions such as cognition [32, 38, 39], memory [32], attention [220], and vigilance [120]. The projection from the prefrontal cortex to the LC may provide tonic activation of the LC [154]. Although the neurotransmitter responsible for this activation is unclear, glutamate may be involved since NMDA and non-NMDA excitatory amino acid receptors are present on LC neurones [170, 380].

3.1.2. Amygdala

The LC receives an input from the central nucleus of the amygdala [54, 55, 261, 332, 384], which thus forms the afferent branch of a reciprocal connection between the LC and the amygdala (see 2.1.3). These reciprocal projections may underlie a role for the LC in processing the emotional valence of stimuli. Complementing the increase in anxiety following LC activation (see 2.1.3), states of anxiety induced by stressful and fear-inducing stimuli, including conditioned fear, are accompanied by increases in LC activity [56, 57, 69, 290, 293] (see 2.2.2, part II) and presumably reflect an increase in the activity of this pathway. Additionally, recent evidence has shown that the administration of anxiogenic drugs of different chemical classes (α2 adrenoceptor antagonist, benzodiazepine inverse agonist, 5HT2C receptor agonist, adenosine receptor antagonist, cholecystokinin analogue) leads to an increase in the expression of c-fos activity in the LC [333], whilst the administration of anxiolytic drugs reduces the activity of neurones in the LC [33]. In addition, the projection from the central nucleus of the amygdala to the LC may also be involved in the observed increase in LC activity in response to stressful stimuli [32]. For example, neurones containing CRF in the central nucleus of the amygdala project to the LC and activate these cells in response to stress [375].

3.1.3. Hypothalamus

3.1.3.1. Ventrolateral Preoptic Area

The GABAergic neurones of the VLPO send an inhibitory projection to the LC [54, 193, 216, 279, 331, 339], which, along with the inhibitory projection from the LC to the VLPO (see 2.2.2.1), forms a reciprocal connection between these two areas (see Figs. (1) and (4) in Part II). During SWS and REM sleep, there is increased release of GABA from the neurones of the VLPO and thus a reduction in the activity of the LC neurones [116, 263]. Indeed, application of GABA to the LC has been found to inhibit cell firing, whilst administration of a GABA receptor antagonist (bicuculline) increases LC activity [116, 219]. In contrast, during wakefulness the inhibitory projection from the LC to the VLPO reduces VLPO neurone activity (see 2.2.2.1) and thus disinhibits the LC from the inhibitory influence of the VLPO [256, 257].

3.1.3.2. Paraventricular Nucleus

There is a well-defined pathway originating in the PVN and projecting to the LC [17, 54, 213, 215, 296, 331, 350, 409]. This PVN projection may form the basis of a second, indirect, pathway to the peripheral autonomic nervous system preganglionic neurones in the brainstem and spinal cord, in addition to the direct projections of the PVN to the preganglionic neurones themselves (see 2.2.2.2). CRF has been suggested as the primary neurotransmitter in the projection to the LC, since excitatory CRF immunoreactive fibres in the PVN have been found to project to, and increase the activity of, the neurones of the LC [296]. Thus, the LC appears to receive CRF inputs from both the paraventricular nucleus and the central nucleus of the amygdala (see 3.1.2).

3.1.3.3. Lateral Hypothalamic/Perifornical Area

The LH/PF densely innervates the neurones of the LC [54, 193, 261] with fibres that contain the orexin peptides [74, 97, 138, 280] (see Fig. (1) in Part II). The orexin system originates solely in the LH/PF, with fibres projecting widely throughout the brainstem and thalamus [78]. Administration of orexin into the LC has been found to increase cell firing [74, 130], suppress REM sleep and increase wakefulness [40]. Along with excitatory projections to other brainstem wakefulness promoting nuclei, for example the TMN [21, 399], the DR [195], and cholinergic neurones of the BF [87] and LDT [45], this excitatory projection to the LC may be involved in the promotion and maintenance of wakefulness [238]. In the sleep disorder narcolepsy, the inability to maintain consistent wakefulness has been related to deficiencies in this orexinergic system [347]. In addition, the orexinergic input to the LC appears to be involved in the maintenance of muscle tone during wakefulness, since orexin microinjections into the LC have been found to facilitate muscle tone [178]. Interestingly, a case report has recently been published describing a patient with a focal lesion in the dorsomedial pontine tegmentum, involving the LC, who developed both narcolepsy and REM sleep behaviour disorder despite normal orexin levels in the cerebrospinal fluid [227]. This report, although only a single case, suggests a key role for the LC in mediating the effects of the orexinergic system on wakefulness and muscle tone.

3.1.3.4. Tuberomamillary Nucleus

The wakefulness-promoting histaminergic neurones of the TMN have been found to project to the LC [149, 194] and histamine H3 receptors have been identified on the cell bodies of LC neurones, where they inhibit noradrenaline release [129, 281]. In contrast to other hypothalamic nuclei, the LC does not appear to project reciprocally to the TMN [95]. The neurones of the TMN are active during wakefulness and quiescent during sleep [129] and this pattern of activity is likely to result from the interaction of the TMN with the VLPO (see Figs. (1) and (4) in Part II). During sleep, when the TMN is quiescent, the inhibitory GABAergic projection from the VLPO is active [324] (see also above). In contrast, during wakefulness when the VLPO is silent (in part due to inhibition from the LC), the TMN is disinhibited and displays a high level of neurone firing [256, 257]. The inhibitory action of the TMN projection to the LC may form part of a negative feedback circuit to restrict the firing of LC neurones during wakefulness: disinhibition of the LC from VLPO neurones may otherwise lead to an ever-increasing rate of LC discharge.

3.1.4. Brainstem

3.1.4.1. Ventral Tegmental Area

Although the LC itself does not contain any dopaminergic cell bodies [240], dopamine-immunoreactive fibres are found within the LC [177, 219] suggesting that there is a dopaminergic projection to this area. Moreover, significant amounts of dopamine [237] and dopaminergic terminals have been found within the LC [331], and both D1-like and D2-like dopamine receptors have been identified on LC neurones [266, 348, 401]. The application of dopamine by reverse microdialysis to the vicinity of the LC inhibits sleep [68] and thus dopamine appears to have an excitatory action on the wakefulness-promoting neurones of the LC.

Neurones have been identified that project from the VTA to the LC [22, 79, 261, 266, 269, 330, 349], an area containing dopaminergic neurones involved in movement, reward, motivation and drug addiction [174, 260, 395]. Indeed, stimulation of the VTA leads to the excitation of the LC-derived noradrenergic dorsal bundle [79] whereas lesions of the VTA result in a reduction in dopamine concentration within the LC [237]. This projection, termed the “mesocoerulear” pathway, may contribute to the maintenance of arousal [305, 306, 308] (see Fig. (1) in Part II). Indeed, the involvement of the VTA in the promotion of wakefulness is supported by evidence indicating that the activation of the VTA produces cortical EEG desynchronisation accompanied by an increase in alertness [77].

3.1.4.2. Raphe Nuclei

It has been reported that the raphe magnus in the CR may project to the LC [328]. This would form a reciprocal connection between the LC and CR for communication relating to the modulation of nociception (see 2.3.2.2). However, there are few studies relating to the outputs of the caudal raphe nuclei. In contrast, there is strong evidence that the serotonergic neurones of the DR project to the LC [54, 176, 215, 261, 276, 329, 331, 382] and this connection is likely to be related to the wakefulness-promoting roles of the two nuclei.

3.1.4.3. Pedunculopontine and Laterodorsal Tegmental Nuclei

The LC is known to receive an input from acetylcholine-releasing neurones [156, 157], and the cholinergic neurones of both the PPT and LDT have been found to project to the LC [164]. As discussed above, the PPT and LDT form two groups of neurones active during either wakefulness or REM sleep [89, 160, 171] and this cholinergic projection to the LC may thus be involved in regulating level of arousal. Perfusions of acetylcholine and cholinoceptor agonists directly into the LC increase the firing of LC neurones [86, 90] and thus increase arousal, suggesting an excitatory role for the PPT and LDT projection to the LC.

Interestingly, the cholinergic projections of the PPT and LDT have also been implicated in the modulation of noradrenergic outputs at terminal regions. Both the noradrenergic projection to the DR [201] and the VLPO [302] are facilitated by a presynaptic cholinergic input to nicotinic receptors on the noradrenergic terminals.

3.1.4.4. Periaqueductal Grey Matter

There is substantial evidence that the LC receives an input from the neurones of the PAG in the midbrain [54, 193, 215, 261, 331], particularly from the dorsolateral cell column of the PAG [49]. The precise functions of the PAG are unclear [174], but recent work has suggested a role in responses to stress, such as the “fight or flight” response, in situations of fear and anxiety [37, 179]. There are also wakefulness-active dopaminergic neurones in the ventral PAG [211] and these may be involved in the activation of LC neurones during wakefulness. In addition the ventral and ventrolateral PAG may be involved in the regulation of sleep-wakefulness state via inhibitory glycinergic projections to the LC [288].

An area in the subcoeruleus surrounding the LC has been identified as containing “REM-on” neurones that receive an inhibitory GABAergic projection from “REM-off” neurones in the ventrolateral PAG [212, 327]. These “REM-on” neurones in turn project to the ventrolateral PAG to inhibit the “REM-off” neurones during REM sleep episodes and may also contain GABA. This reciprocal connection has been termed a “flip-flop switch” and it has been suggested that this connection may underlie the transition into REM sleep [212, 327].

3.1.4.5. Medulla

Two groups of neurones in the rostral medulla have major inputs to the LC: the PrH [17, 93, 94, 215, 331] and the RVLM [13] (also described as the PGi) [17, 128, 215, 331, 343]. The projection from the PrH to the LC contains GABAergic neurones and is thus inhibitory to LC neurone activity [93, 94]. This GABAergic projection is likely to be involved in the inhibition of LC activity during REM sleep [381]. Indeed, electrical stimulation of GABAergic neurones in the PrH increases REM sleep duration [169] and it is likely that this occurs via the inhibitory projection to the wakefulness-promoting neurones of the LC. In contrast, the projection from the RVLM excites LC neurone activity via the release of glutamate [331] and electrical stimulation of the RVLM increases the activity within the LC [91, 92]. The projection from the RVLM to the LC is likely to be involved in the modulation of autonomic functioning, since the RVLM is known to be centrally involved in the regulation of cardiovascular function (see 2.3.2.1), and may provide an integrated input to the LC regarding this information [262, 376]. Interestingly, clonidine microinjection in the RVLM resulted in sedation in rats, suggesting that the stimulation of inhibitory α2-adrenoceptors on the RVLM removes the excitatory input to the LC [400]. The control of cardiovascular function and arousal are thus intricately related.

In addition to the GABAergic and glutamatergic projections to the LC from the rostral medulla, both the PrH and the PGi innervate the LC with fibres containing the endogenous opiate enkephalin [83, 155, 372]. These projections activate opiate receptors found in high concentrations in the LC to inhibit cell firing [372, 364] and the administration of endogenous opioids or opiate agonists reduces LC spontaneous firing [146, 183, 278, 371]. The LC may thus be involved in opiate-induced analgesia [83] and opiate withdrawal [155].

The adrenergic neurones of the ventrolateral medulla (cell groups C1-C3) also project to the LC [261], and likely to contribute to the role of the LC in cardiovascular regulation (see Section 1.2 in Part II).

3.1.5. Cerebellum

There is some evidence that the nuclei of the cerebellum innervate the LC [261]. However, several reports describing the efferent projections from these nuclei do not describe terminals in the LC (for example, 190, 250, 253).

3.1.6. Spinal Cord

In addition to receiving a dense innervation from the LC, the dorsal horn has been found to project to the LC in return [54, 67, 261]. It has been suggested that this pathway may communicate information relating to the detection of nociceptive and/or thermal stimuli from sensory spinal nuclei [67, 261].

3.2. Modulation by Autoreceptors

There are α2-adrenoceptors located presynaptically on LC neurones which act to inhibit the activity of these neurones [319, 358, 404] and it has been suggested that noradrenaline release from the LC is under tonic inhibitory control via these autoreceptors [286, 331]. Indeed, stimulation of the autoreceptors by noradrenaline application reduces LC firing rate [85, 146, 393, 394], whilst blockade of the autoreceptors increases LC activity [331] and potentiates the activity evoked by glutamate administration [147, 331].

The autoreceptors are endogenously activated by the release of noradrenaline from LC collaterals and thus provide a self-regulating mechanism of negative feedback [4, 85]. The autoreceptors may also be activated by noradrenaline escaping from the dendrites of LC neurones [99, 142, 285] and also by adrenaline released from an adrenergic innervation arising in the PGi of the rostral medulla [372]. Interestingly, the sleep disorder narcolepsy has been associated with an increase in autoreceptors within the LC, suggesting that an increase in LC inhibition may contribute to this disorder [109].

In addition, it has been demonstrated that μ-opiate receptors are co-localised with α2-adrenoceptor autoreceptors on LC neurones [5, 372] and utilise the same ion channel mediating an increase in potassium conductance [102]. Thus, these receptors also act to inhibit neurone activity within the LC (see 3.1.1.6, Part II).

CONCLUSIONS

In conclusion, it is clear that the LC is a major noradrenergic nucleus, giving rise to fibres innervating most structures of the neuraxis in a highly specific manner. These structures control a number of physiological processes including the regulation of arousal and autonomic function and the LC is, therefore, central to the regulation of these processes. The LC is known to be a major wakefulness-promoting nucleus, with activation of the LC resulting in an increase in EEG signs of alertness. This alerting effect of LC activation results from dense excitatory projections to the majority of the cerebral cortex, wakefulness-promoting cholinergic neurones of the basal forebrain, cortically-projecting excitatory neurones of the thalamus, wakefulness-promoting serotonergic neurones of the dorsal raphe, wakefulness-promoting cholinergic neurones of the pedunculopontine tegmental nucleus and laterodorsal tegmental nucleus, and substantial inhibitory projections to sleep-promoting GABAergic neurones of the basal forebrain and ventrolateral preoptic area. It is also clear that the LC plays an important role in controlling autonomic function, where LC activation produces an increase in sympathetic activity and a concomitant decrease in parasympathetic activity. This contribution to the control of autonomic activity results both from the direct projections to the sympathetic and parasympathetic divisions of the spinal cord and from the indirect projections to various nuclei influencing the autonomic system, including the parasympathetic dorsal nucleus of the vagus and nucleus ambiguus and the sympathetic rostroventrolateral medulla, involved in cardiovascular regulation, the parasympathetic Edinger-Westphal nucleus, involved in pupil constriction, the caudal raphe, the salivatory nuclei, the paraventricular nucleus, and the amygdala. The control of arousal and autonomic function is thus inseparably linked, largely via the involvement of the LC. Changes in LC activity result in complex patterns of neuronal activity throughout the brain since the noradrenergic outputs from the LC can exert both excitatory effects via α1-adrenoceptors and inhibitory effects via α2-adrenoceptors. The effect of LC activation on arousal and autonomic function is therefore both interrelated and intricate, based on the compound effects of multiple projections to areas of influence.

Abbreviations

5HT

5-hydroxytryptamine

BF

Basal forebrain

CR

Caudal raphe

CRF

Corticotrophin-releasing factor

DMV

Dorsal motor nucleus of the vagus

DR

Dorsal raphe

EEG

Electroencephalogram

EMG

Electromyogram

EWN

Edinger-Westphal nucleus

GABA

Gamma-aminobutyric acid

GH

Growth hormone

GHRH

Growth hormone releasing hormone

IML

Intermediolateral cell column

LC

Locus coeruleus

LDT

Laterodorsal tegmental nucleus

LH/PF

Lateral hypothalamus and perifornical area

NMDA

N-methyl D-aspartate

PAG

Periaqueductal grey matter

PGi

Nucleus paragigantocellularis lateralis

PPT

Pedunculopontine tegmental nucleus

PrH

Nucleus prepositus hypoglossi

PVN

Paraventricular nucleus

REM

Rapid eye movement

RVLM

Rostroventrolateral medulla

SWS

Slow wave sleep

TMN

Tuberomamillary nucleus

VLPO

Ventrolateral preoptic area

VTA

Ventral tegmental area

REFERENCES

  • 1.Abduljawad KAJ, Langley RW, Bradshaw CM, Szabadi E. Effects of clonidine and diazepam on the acoustic startle response and on its inhibition by ‘prepulses’ in man. J. Psychopharmacol. 1997;11:29–34. doi: 10.1177/026988119701100110. [DOI] [PubMed] [Google Scholar]
  • 2.Abduljawad KAJ, Langley RW, Bradshaw CM, Szabadi E. Effects of clonidine and diazepam on prepulse inhibition of the acoustic startle response and the N1/P2 auditory evoked potential in man. J. Psychopharmacol. 2001;15:237–242. doi: 10.1177/026988110101500402. [DOI] [PubMed] [Google Scholar]
  • 3.Acosta-Martinez M, Fiber JM, Brown RD, Etgen AM. Localization of α1B-adrenergic receptor in female rat brain regions involved in stress and neuroendocrine function. Neurochem. Int. 1999;35:383–391. doi: 10.1016/s0197-0186(99)00077-7. [DOI] [PubMed] [Google Scholar]
  • 4.Aghajanian GK, Cedarbaum JM, Wang RY. Evidence for norepinephrine-mediated collateral inhibition of locus coeruleus neurons. Brain Res. 1977;136:570–577. doi: 10.1016/0006-8993(77)90083-x. [DOI] [PubMed] [Google Scholar]
  • 5.Aghajanian GK, Wang YY. Common alpha-1 and opiate effector mechanisms in the locus coeruleus intracellular studies in brain slices. Neuropharmacology. 1987;26:793–799. doi: 10.1016/0028-3908(87)90054-2. [DOI] [PubMed] [Google Scholar]
  • 6.Aldes LD. Topographically organized projections from the nucleus subcoeruleus to the hypoglossal nucleus in the rat: A light and electron microscopic study with complementary axonal transport techniques. J. Comp. Neurol. 1990;302:643–656. doi: 10.1002/cne.903020318. [DOI] [PubMed] [Google Scholar]
  • 7.Allen GV, Cechetto DF. Serotoninergic and nonserotoninergic neurons in the medullary raphe system have axon collateral projections to autonomic and somatic cell groups in the medulla and spinal cord. J. Comp. Neurol. 1994;350:357–366. doi: 10.1002/cne.903500303. [DOI] [PubMed] [Google Scholar]
  • 8.Amaral DG, Sinnamon HM. The locus coeruleus: neurobiology of a central noradrenergic nucleus. Prog. Neurobiol. 1977;9:147–196. doi: 10.1016/0301-0082(77)90016-8. [DOI] [PubMed] [Google Scholar]
  • 9.Andrews GD, Lavin A. Methylphenidate increases cortical excitability via activation of alpha-2 noradrenergic receptors. Neuropsychopharmacology. 2006;31:594–601. doi: 10.1038/sj.npp.1300818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Anselmo-Franci JA, Peres-Polon VL, da Rocha-Barros VM, Moreira ER, Franci CR, Rocha MJA. C-fos expression and electrolytic lesions studies reveal activation of the posterior region of locus coeruleus during hemorrhage induced hypotension. Brain Res. 1998;799:278–284. doi: 10.1016/s0006-8993(98)00468-5. [DOI] [PubMed] [Google Scholar]
  • 11.Appel NM, Elde RP. The intermediolateral cell column of the thoracic spinal cord is comprised of target-specific subnuclei: evidence from retrograde transport studies and immunohistochemistry. J. Neurosci. 1988;8:1767–1775. doi: 10.1523/JNEUROSCI.08-05-01767.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Arnsten AF, Goldman-Rakic PS. Selective prefrontal cortical projections to the region of the locus coeruleus and raphe nuclei in rhesus monkey. Brain Res. 1984;306:9–18. doi: 10.1016/0006-8993(84)90351-2. [DOI] [PubMed] [Google Scholar]
  • 13.Astier B, Van Bockstaele EJ, Aston-Jones G, Pieribone VA. Anatomical evidence for multiple pathways leading from the rostral ventrolateral medulla (nucleus paragigantocellularis) to the locus coeruleus in the rat. Neurosci. Lett. 1990;118:141–146. doi: 10.1016/0304-3940(90)90612-d. [DOI] [PubMed] [Google Scholar]
  • 14.Aston-Jones G. Brain structures and receptors involved in alertness. Sleep Med. 2005;6(suppl. 1):S3–S7. doi: 10.1016/s1389-9457(05)80002-4. [DOI] [PubMed] [Google Scholar]
  • 15.Aston-Jones G, Bloom FE. Activity of norepinephrine-containing locus coeruleus neurons in behaving rats anticipates fluctuations in the sleep- waking cycle. J. Neurosci. 1981;1:876–886. doi: 10.1523/JNEUROSCI.01-08-00876.1981. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Aston-Jones G, Chiang C, Alexinsky T. Discharge of noradrenergic locus coeruleus neurons in behaving rats and monkeys suggests a role in vigilance. Prog. Brain Res. 1991;88:501–520. doi: 10.1016/s0079-6123(08)63830-3. [DOI] [PubMed] [Google Scholar]
  • 17.Aston-Jones G, Ennis M, Pieribone VA, Nickell WT, Shipley MT. The brain nucleus locus coeruleus: restricted afferent control of a broad efferent network. Science. 1986;234:734–737. doi: 10.1126/science.3775363. [DOI] [PubMed] [Google Scholar]
  • 18.Baldo BA, Daniel RA, Berridge CW, Kelley AE. Overlapping distributions of orexin/hypocretin- and dopamine-beta-hydroxylase immunoreactive fibers in rat brain regions mediating arousal, motivation, and stress. J. Comp. Neurol. 2003;464:220–237. doi: 10.1002/cne.10783. [DOI] [PubMed] [Google Scholar]
  • 19.Barman SM, Gebber GL. Axonal projection patterns of ventrolateral medullospinal sympathoexcitatory neurons. J. Neurophysiol. 1985;53:1551–1566. doi: 10.1152/jn.1985.53.6.1551. [DOI] [PubMed] [Google Scholar]
  • 20.Bay KD, Mamiya K, Good CH, Skinner RD, Garcia-Rill E. Alpha-2 adrenergic regulation of pedunculopontine nucleus neurons during development. Neuroscience. 2006;141:769–779. doi: 10.1016/j.neuroscience.2006.04.045. [DOI] [PubMed] [Google Scholar]
  • 21.Bayer L, Eggermann E, Serafin M, Saint-Mleux B, Machard D, Jones B, Mϋhlethaler M. Orexins (hypocretins) directly excite tuberomammillary neurons. Eur. J. Neurosci. 2001;14:1571–1575. doi: 10.1046/j.0953-816x.2001.01777.x. [DOI] [PubMed] [Google Scholar]
  • 22.Beckstead RM, Domesick VB, Nauta WJ. Efferent connections of the substantia nigra and ventral tegmental area in the rat. Brain Res. 1979;175:191–217. doi: 10.1016/0006-8993(79)91001-1. [DOI] [PubMed] [Google Scholar]
  • 23.Ben-Jonathan N. Dopamine: a prolactin-inhibiting hormone. Endocr. Rev. 1985;6:564–589. doi: 10.1210/edrv-6-4-564. [DOI] [PubMed] [Google Scholar]
  • 24.Berridge CW, Abercrombie ED. Relationship between locus coeruleus discharge rates and rates of norepinephrine release within neocortex as assessed by in vivo microdialysis. Neuroscience. 1999;93:1263–1270. doi: 10.1016/s0306-4522(99)00276-6. [DOI] [PubMed] [Google Scholar]
  • 25.Berridge CW, Foote SL. Effects of locus coeruleus activation on electroencephalographic activity in neocortex and hippocampus. J. Neurosci. 1991;11:3135–3145. doi: 10.1523/JNEUROSCI.11-10-03135.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Berridge CW, Foote SL. Enhancement of behavioural and electroencephalographic indices of waking following stimulation of noradrenergic β-receptors within the medial septal region of the basal forebrain. J. Neurosci. 1996;16:6999–7009. doi: 10.1523/JNEUROSCI.16-21-06999.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Berridge CW, Isaac SO, España RA. Additive wake-promoting actions of medial basal forebrain noradrenergic α1- and β-receptor stimulation. Behav. Neurosci. 2003;117:350–359. doi: 10.1037/0735-7044.117.2.350. [DOI] [PubMed] [Google Scholar]
  • 28.Berridge CW, O’Neill J. Differential sensitivity to the wake- promoting effects of norepinephrine within the medial preoptic area and the substantia innominata. Behav. Neurosci. 2001;115:165–174. doi: 10.1037/0735-7044.115.1.165. [DOI] [PubMed] [Google Scholar]
  • 29.Berridge CW, O’Neill J, Wifler K. Amphetamine acts within the medial basal forebrain to initiate and maintain alert waking. Neuroscience. 1999;93:885–896. doi: 10.1016/s0306-4522(99)00271-7. [DOI] [PubMed] [Google Scholar]
  • 30.Berridge CW, Page ME, Valentino RJ, Foote SL. Effects of locus coeruleus activation on electroencephalographic activity in neocortex and hippocampus. Neuroscience. 1993;55:381–393. doi: 10.1016/0306-4522(93)90507-c. [DOI] [PubMed] [Google Scholar]
  • 31.Berridge CW, Stellick RL, Schmeichel BE. Wake-promoting actions of medial basal forebrain β2 receptor stimulation. Behav. Neurosci. 2005;119:743–751. doi: 10.1037/0735-7044.119.3.743. [DOI] [PubMed] [Google Scholar]
  • 32.Berridge CW, Waterhouse BD. The locus coeruleus-noradrenergic system: modulation of behavioral state and state-dependent cognitive processes. Brain Res. Rev. 2003;42:33–84. doi: 10.1016/s0165-0173(03)00143-7. [DOI] [PubMed] [Google Scholar]
  • 33.Berrocoso E, Micó JA, Ugedo L. In vivo effect of tramadol on locus coeruleus neurons is mediated by α2-adrenoceptors and modulated by serotonin. Neuropharmacology. 2006;51:146–153. doi: 10.1016/j.neuropharm.2006.03.013. [DOI] [PubMed] [Google Scholar]
  • 34.Bie B, Fields HL, Williams JT, Pan ZZ. Roles of α1- and α2- adrenoceptors in the nucleus raphe magnus in opioid analgesia and opioid abstinence-induced hyperalgesia. J. Neurosci. 2003;23:7950–7957. doi: 10.1523/JNEUROSCI.23-21-07950.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Blair ML, Piekut D, Want A, Olschowka JA. Role of the hypothalamic paraventricular nucleus in cardiovascular regulation. Clin. Exp. Pharmacol. Physiol. 1996;23:161–165. doi: 10.1111/j.1440-1681.1996.tb02590.x. [DOI] [PubMed] [Google Scholar]
  • 36.Bonaz B, Taché Y. Induction of Fos immunoreactivity in the rat brain after cold-restraint induced gastric lesions and fecal excretion. Brain Res. 1994;652:56–64. doi: 10.1016/0006-8993(94)90316-6. [DOI] [PubMed] [Google Scholar]
  • 37.Borelli KG, Ferreira-Netto C, Coimbra NC, Brandao ML. Fos- like immunoreactivity in the brain associated with freezing or escape induced by inhibition of either glutamic acid decarboxylase or GABAA receptors in the dorsal periaqueductal gray. Brain Res. 2005;1051:100–111. doi: 10.1016/j.brainres.2005.05.068. [DOI] [PubMed] [Google Scholar]
  • 38.Bouret S, Sara SJ. Reward expectation, orientation of attention and locus coeruleus-medial frontal cortex interplay during learning. Eur. J. Neurosci. 2004;20:791–802. doi: 10.1111/j.1460-9568.2004.03526.x. [DOI] [PubMed] [Google Scholar]
  • 39.Bouret S, Sara SJ. Network reset: a simplified overarching theory of locus coeruleus noradrenaline function. Trends Neurosci. 2005;28:574–582. doi: 10.1016/j.tins.2005.09.002. [DOI] [PubMed] [Google Scholar]
  • 40.Bourgin P, Huitrón-Reséndiz S, Spier AD, Fabre V, Morte B, Criado JR, Sutcliffe JG, Henriksen SJ, de Lecea L. Hypocretin-1 modulates rapid eye movement sleep through activation of locus coeruleus neurons. J. Neurosci. 2000;20:7760–7765. doi: 10.1523/JNEUROSCI.20-20-07760.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Bradley RM, Mistretta CM, Bates CA, Killackey HP. Transganglionic transport of HRP from the circumvallate papilla of the rat. Brain Res. 1985;361:154–161. doi: 10.1016/0006-8993(85)91285-5. [DOI] [PubMed] [Google Scholar]
  • 42.Breen LA, Burde RM, Loewy AD. Brainstem connections to the Edinger-Westphal nucleus of the cat: a retrograde tracer study. Brain Res. 1983;261:303–306. doi: 10.1016/0006-8993(83)90633-9. [DOI] [PubMed] [Google Scholar]
  • 43.Brown RE, Sergeeva OA, Eriksson KS, Haas HL. Convergent excitation of dorsal raphe serotonin neurons by multiple arousal systems (orexin/hypocretin, histamine and noradrenaline) J. Neurosci. 2002;22:8850–8859. doi: 10.1523/JNEUROSCI.22-20-08850.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Buijs RM, la Fleur SE, Wortel J, van Heyningen C, Zuiddam L, Mettenleiter TC, Kalsbeek A, Nagai K, Nijima A. The suprachiasmatic nucleus balances sympathetic and parasympathetic output to peripheral organs through separate preautonomic neurons. J. Comp. Neurol. 2003;464:36–48. doi: 10.1002/cne.10765. [DOI] [PubMed] [Google Scholar]
  • 45.Burlet S, Tyler CJ, Leonard CS. Direct and indirect excitation of laterodorsal tegmental neurons by hypocretin/orexin peptides: implications for wakefulness and narcolepsy. J. Neurosci. 2002;22:2862–2872. doi: 10.1523/JNEUROSCI.22-07-02862.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Byrum CE, Guyenet PG. Afferent and efferent connections of the A5 noradrenergic cell group in the rat. J. Comp. Neurol. 1987;261:529–542. doi: 10.1002/cne.902610406. [DOI] [PubMed] [Google Scholar]
  • 47.Cahusac PMB, Morris R, Hill RG. A pharmacological study of the modulation of neuronal and behavioural nociceptive responses in the rat trigeminal region. Brain Res. 1995;700:70–82. doi: 10.1016/0006-8993(95)00927-i. [DOI] [PubMed] [Google Scholar]
  • 48.Calaresu FR, Yardley CP. Medullary basal sympathetic tone. Ann. Rev. Physiol. 1988;50:511–524. doi: 10.1146/annurev.ph.50.030188.002455. [DOI] [PubMed] [Google Scholar]
  • 49.Cameron AA, Khan IA, Westlund KN, Willis WD. The efferent projections of the periaqueductal gray in the rat: a Phaseolus vulgaris- leucoagglutinin study. II Descending projections. J. Comp. Neurol. 1995;351:585– 601. doi: 10.1002/cne.903510408. [DOI] [PubMed] [Google Scholar]
  • 50.Cape EG, Jones BE. Differential modulation of high-frequency γ- electroencephalogram activity and sleep-wake state by noradrenaline and serotonin microinjections into the region of cholinergic basalis neurons. J. Neurosci. 1998;18:2653–2666. doi: 10.1523/JNEUROSCI.18-07-02653.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Capuano CA, Leibowitz SF, Barr GA. The pharmaco-ontogeny of the perifornical lateral hypothalamic beta 2-adrenergic and dopaminergic receptor systems mediating epinephrine- and dopamine-induced suppression of feeding in the rat. Brain Res. Dev. Brain Res. 1992;70:1–7. doi: 10.1016/0165-3806(92)90098-h. [DOI] [PubMed] [Google Scholar]
  • 52.Carpenter MB, Periera AB, Guha N. Immunocytochemistry of oculomotor afferents in the squirrel monkey (Saimiri sciureus) J. Hirnforsch. 1992;33:151–167. [PubMed] [Google Scholar]
  • 53.Caverson MM, Ciriello J, Calaresu FR. Direct pathway from cardiovascular neurons in the ventrolateral medulla to the region of the intermediolateral nucleus of the upper thoracic cord: an anatomical and electrophysiological investigation in the cat. J. Auton. Nerv. Syst. 1983;9:451–475. doi: 10.1016/0165-1838(83)90007-3. [DOI] [PubMed] [Google Scholar]
  • 54.Cedarbaum JM, Aghajanian GK. Afferent projections to the rat locus coeruleus as determined by a retrograde tracing technique. J. Comp. Neurol. 1978;178:1–16. doi: 10.1002/cne.901780102. [DOI] [PubMed] [Google Scholar]
  • 55.Charney DS, Grillon C, Bremner JD. The neurobiological basis of anxiety and fear: circuits, mechanisms, and neurochemical interactions (part 1) Neuroscientist. 1998;4:35–44. [Google Scholar]
  • 56.Charney DS, Redmond DE Jr. Neurobiological mechanisms in human anxiety. Neuropharmacology. 1983;22:1531–1536. doi: 10.1016/0028-3908(83)90122-3. [DOI] [PubMed] [Google Scholar]
  • 57.Charney DS, Woods SW, Nagy LM, Southwick SM, Krystal JH, Heninger GR. Noradrenergic function in panic disorder. J. Clin. Psychiatry. 1990;51 (Suppl A):5–11. [PubMed] [Google Scholar]
  • 58.Chen SY, Chai CY. Coexistence of neurons integrating urinary bladder activity and pelvic nerve activity in the same cardiovascular areas of the pontomedulla in cats. Chin. J. Physiol. 2002;45:41–50. [PubMed] [Google Scholar]
  • 59.Chen F-J, Sara SJ. Locus coeruleus activation by foot shock or electrical stimulation inhibits amygdala neurons. Neuroscience. 2007;144:472–481. doi: 10.1016/j.neuroscience.2006.09.037. [DOI] [PubMed] [Google Scholar]
  • 60.Cheng Z, Zhang H, Yu J, Wurster RD, Gozal D. Attenuation of baroreflex sensitivity after domoic acid lesion of the nucleus ambiguus of rats. J. Appl. Physiol. 2004;96:1137–1145. doi: 10.1152/japplphysiol.00391.2003. [DOI] [PubMed] [Google Scholar]
  • 61.Chibuzo GA, Cummings JF. Motor and sensory centers for the innervation of mandibular and sublingual salivary glands: a horseradish peroxidase study in the dog. Brain Res. 1980;189:301–313. doi: 10.1016/0006-8993(80)90092-x. [DOI] [PubMed] [Google Scholar]
  • 62.Chou TC, Bjorkum AA, Gaus SE, Lu J, Scammell TE, Saper CB. Afferents to the ventrolateral preoptic nucleus. J. Neurosci. 2002;22:977–990. doi: 10.1523/JNEUROSCI.22-03-00977.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Clark FM, Proudfit HK. The projection of locus coeruleus neurons to the spinal cord in the rat determined by anterograde tracing combined with immunocytochemistry. Brain Res. 1991;538:231–245. doi: 10.1016/0006-8993(91)90435-x. [DOI] [PubMed] [Google Scholar]
  • 64.Conway S, Richardson L, Speciale S, Moherek R, Mauceri H, Krulich L. Interaction between norepinephrine and serotonin in the neuroendocrine control of growth hormone release in the rat. Endocrinology. 1990;126:1022–1030. doi: 10.1210/endo-126-2-1022. [DOI] [PubMed] [Google Scholar]
  • 65.Coote JH. Noradrenergic projections to the spinal cord and their role in cardiovascular control. J. Auton Nerv. Syst. 1985;14:255–262. doi: 10.1016/0165-1838(85)90114-6. [DOI] [PubMed] [Google Scholar]
  • 66.Couto LB, Moroni CR, dos Reis Ferreira CM, Elias-Filho DH, Parada CA, Pelá IR, Coimbra NC. Descriptive and functional neuroanatomy of locus coeruleus-noradrenaline-containing neurons involvement in bradykinin-induced antinociception on principle sensory trigeminal nucleus. J. Chem. Neuroanat. 2006;32:28–45. doi: 10.1016/j.jchemneu.2006.03.003. [DOI] [PubMed] [Google Scholar]
  • 67.Craig AD. Spinal and trigeminal lamina I input to the locus coeruleus anterogradely labeled with Phaseolus vulgaris leucoagglutinin (PHA-L) in the cat and monkey. Brain Res. 1992;584:325–328. doi: 10.1016/0006-8993(92)90915-v. [DOI] [PubMed] [Google Scholar]
  • 68.Crochet S, Sakai K. Dopaminergic modulation of behavioural states in mesopontine tegmentum: A reverse microdialysis study in freely moving cats. Sleep. 2003;26:801–806. doi: 10.1093/sleep/26.7.801. [DOI] [PubMed] [Google Scholar]
  • 69.Cullinan WE, Herman JP, Battaglia DF, Akil H, Watson SJ. Pattern and time course of immediate early gene expression in rat brain following acute stress. Neuroscience. 1995;64:477–505. doi: 10.1016/0306-4522(94)00355-9. [DOI] [PubMed] [Google Scholar]
  • 70.Dahlström A, Fuxe K. Evidence for the existence of monoamine- containing neurons in the central nervous system. I. Demonstration of monoamines in the cell bodies of brainstem neurons. Acta Physiol. Scand. Suppl. 1964;232:1–55. [PubMed] [Google Scholar]
  • 71.Damasio AR. Emotion in the perspective of an integrated nervous system. Brain Res. Rev. 1998;26:83–86. doi: 10.1016/s0165-0173(97)00064-7. [DOI] [PubMed] [Google Scholar]
  • 72.Dampney RAL. Functional organization of central pathways regulating the cardiovascular system. Physiol. Rev. 1994;74:332–364. doi: 10.1152/physrev.1994.74.2.323. [DOI] [PubMed] [Google Scholar]
  • 73.Danysz W, Dyr W, Plaznik A, Kostowski W. The effect of microinjections of clonidine into the locus coeruleus on cortical EEG in rats. Pol. J. Pharmacol. Pharm. 1989;41:45–50. [PubMed] [Google Scholar]
  • 74.Date Y, Ueta Y, Yamashita H, Yamaguchi H, Matsukura S, Kangawa K, Sakurai T, Yanahisawa M, Nakazato M. Orexins, orexigenic hypothalamic peptides, interact with autonomic, neuroendocrine and neuroregulatory systems. Proc. Natl. Acad. Sci. USA. 1999;96:748–753. doi: 10.1073/pnas.96.2.748. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Day HEW, Campeau S, Watson SJ Jr, Akil H. Distribution of α1a-, α1b- and α1d-adrenergic receptor mRNA in the rat brain and spinal cord. J. Chem. Neuroanat. 1997;13:115–139. doi: 10.1016/s0891-0618(97)00042-2. [DOI] [PubMed] [Google Scholar]
  • 76.Day HEW, Greenwood BN, Hammack SE, Watkins LR, Fleshner M, Maier SF, Campeau S. Differential expression of 5HT-1A, α1b adrenergic, CRF-R1, and CRF-R2 receptor mRNA in serotonergic, γ- aminobutyric acidergic, and catecholaminergic cells of the rat dorsal raphe nucleus. J. Comp. Neurol. 2004;474:364–378. doi: 10.1002/cne.20138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.De Keyser J, Ebinger G, Vauquelin G. Evidence for a widespread dopaminergic innervation of the human cerebral neocortex. Neurosci. Lett. 1989;104:281–285. doi: 10.1016/0304-3940(89)90589-2. [DOI] [PubMed] [Google Scholar]
  • 78.de Lecea L, Kilduff TS, Peyron C, Gao X-B, Foye PE, Danielson PE, Fukuhara C, Battenberg ELF, Gautvik VT, Bartlett FS II, Frankel WN, van den Pol AN, Bloom FE, Gautvik KM, Sutcliffe JG. The hypocretins: hypothalamus-specific peptides with neuroexcitatory activity. Proc. Natl. Acad. Sci. USA. 1998;95:322–327. doi: 10.1073/pnas.95.1.322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Deutch AY, Goldstein M, Roth RH. Activation of the locus coeruleus induced by selective stimulation of the ventral tegmental area. Brain Res. 1986;363:307–314. doi: 10.1016/0006-8993(86)91016-4. [DOI] [PubMed] [Google Scholar]
  • 80.Devauges V, Sara SJ. Memory retrieval enhancement by locus coeruleus stimulation: evidence for mediation by beta-receptors. Behav. Brain Res. 1991;43:93–97. doi: 10.1016/s0166-4328(05)80056-7. [DOI] [PubMed] [Google Scholar]
  • 81.Domyancic AV, Morilak DA. Distribution of α1A adrenergic receptor mRNA in the rat brain visualized by in situ hybridization. J. Comp. Neurol. 1997;386:358–378. doi: 10.1002/(sici)1096-9861(19970929)386:3<358::aid-cne3>3.0.co;2-0. [DOI] [PubMed] [Google Scholar]
  • 82.Drolet G, Gauthier P. Peripheral and central mechanisms of the pressor response elicited by stimulation of the locus coeruleus in the rat. Can. J. Physiol. Pharmacol. 1985;63:599–605. doi: 10.1139/y85-100. [DOI] [PubMed] [Google Scholar]
  • 83.Drolet G, Van Bockstaele EJ, Aston-Jones G. Robust enkephalin innervation of the locus coeruleus from the rostral medulla. J. Neurosci. 1992;12:3162–3174. doi: 10.1523/JNEUROSCI.12-08-03162.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Ebert U. Noradrenalin enhances the activity of cochlear nucleus neurons in the rat. Eur. J. Neurosci. 1996;8:1306–1314. doi: 10.1111/j.1460-9568.1996.tb01299.x. [DOI] [PubMed] [Google Scholar]
  • 85.Egan TM, Henderson G, North RA, Williams JT. Noradrenaline- mediated synaptic inhibition in rat locus coeruleus neurones. J. Physiol. 1983;345:477–488. doi: 10.1113/jphysiol.1983.sp014990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Egan TM, North RA. Acetylcholine acts on m2-muscarinic receptors to excite rat locus coeruleus neurones. Br. J. Pharmacol. 1985;85:733–735. doi: 10.1111/j.1476-5381.1985.tb11070.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Eggermann E, Serafin M, Bayer L, Machard D, Saint-Mleux B, Jones BE, Mϋhlethaler M. Orexins/hypocretins excite basal forebrain cholinergic neurones. Neuroscience. 2001;108:177–181. doi: 10.1016/s0306-4522(01)00512-7. [DOI] [PubMed] [Google Scholar]
  • 88.Eisenach JC, De Kock M, Klimscha W. Alpha(2)-adrenergic agonists for regional anesthesia. A clinical review of clonidine (198401995) Anesthesiology. 1996;85:655–674. doi: 10.1097/00000542-199609000-00026. [DOI] [PubMed] [Google Scholar]
  • 89.el Mansari M, Sakai K, Jouvet M. Unitary characteristics of presumptive cholinergic tegmental neurons during the sleep-waking cycle in freely moving cats. Exp. Brain Res. 1989;76:519–529. doi: 10.1007/BF00248908. [DOI] [PubMed] [Google Scholar]
  • 90.Engberg G, Svensson TH. Pharmacological analysis of a cholinergic receptor mediated regulation of brain norepinephrine neurons. J. Neural Transm. 1980;49:137–150. doi: 10.1007/BF01245220. [DOI] [PubMed] [Google Scholar]
  • 91.Ennis M, Aston-Jones G. A potent excitatory input to the nucleus locus coeruleus from the ventrolateral medulla. Neurosci. Lett. 1986;71:299–305. doi: 10.1016/0304-3940(86)90637-3. [DOI] [PubMed] [Google Scholar]
  • 92.Ennis M, Aston-Jones G. Activation of locus coeruleus from nucleus paragigantocellularis: a new excitatory amino acid pathway in brain. J. Neurosci. 1988;8:3644–3657. doi: 10.1523/JNEUROSCI.08-10-03644.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Ennis M, Aston-Jones G. GABA-mediated inhibition of locus coeruleus from the dorsomedial rostral medulla. J. Neurosci. 1989;9:2973–2981. doi: 10.1523/JNEUROSCI.09-08-02973.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Ennis M, Aston-Jones G. Potent inhibitory input to locus coeruleus from the nucleus prepositus hypoglossi. Brain Res. Bull. 1989;22:793–803. doi: 10.1016/0361-9230(89)90022-1. [DOI] [PubMed] [Google Scholar]
  • 95.Ericson H, Blomqvist A, Kohler C. Brainstem afferents to the Tuberomammillary nucleus in the rat brain with special reference to monoaminergic innervation. J. Comp. Neurol. 1989;281:169–192. doi: 10.1002/cne.902810203. [DOI] [PubMed] [Google Scholar]
  • 96.España RA, Berridge CW. Organisation of noradrenergic efferents to arousal-related basal forebrain structures. J. Comp. Neurol. 2006;496:668–683. doi: 10.1002/cne.20946. [DOI] [PubMed] [Google Scholar]
  • 97.España RA, Reis KM, Valentino RJ, Berridge CW. Organization of hypocretin/orexin efferents to locus coeruleus and basal forebrain arousal-related structures. J. Comp. Neurol. 2005;481:160–178. doi: 10.1002/cne.20369. [DOI] [PubMed] [Google Scholar]
  • 98.Fallon JH, Koziell DA, Moore RY. Catecholamine innervation of the basal forebrain. II. Amygdala, suprarhinal cortex and entorhinal cortex. J. Comp. Neurol. 1978;180:509–532. doi: 10.1002/cne.901800308. [DOI] [PubMed] [Google Scholar]
  • 99.Fernández-Pastor B, Mateo Y, Gómez-Urquijo S, Meana JJ. Characterization of neurotransmitter release in the locus coeruleus of freely moving awake rats by in vivo microdialysis. Psychopharmacology (Berl.) 2005;180:570–579. doi: 10.1007/s00213-005-2181-y. [DOI] [PubMed] [Google Scholar]
  • 100.Ferry B, Roozendaal B, McGaugh JL. Basolateral amygdala noradrenergic influences on memory storage are mediated by an interaction between β- and α1-adrenoceptors. J. Neurosci. 1999;19:5119–5123. doi: 10.1523/JNEUROSCI.19-12-05119.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Ferry B, Roozendaal B, McGaugh JL. Involvement of α1- adrenoceptors in the basolateral amygdala in modulation of memory storage. Eur. J. Pharmacol. 1999;372:9–16. doi: 10.1016/s0014-2999(99)00169-7. [DOI] [PubMed] [Google Scholar]
  • 102.Foote SL, Aston-Jones GS. Pharmacology and physiology of central noradrenergic systems Psychopharmacology: The Fourth Generation of Progress. In: Bloom FE, Kupfer DJ, editors. The Fourth Generation of Progress. New York: Raven Press; pp. 335–345. [Google Scholar]
  • 103.Foote SL, Aston-Jones G, Bloom FE. Impulse activity of locus coeruleus neurons in awake rats and monkeys is a function of sensory stimulation and arousal. Proc. Natl. Acad. Sci. USA. 1980;77:3033–3037. doi: 10.1073/pnas.77.5.3033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Foote SL, Berridge CW, Adams LM, Pineda JA. Electrophysiological evidence for the involvement of the locus coeruleus in alerting, orienting, and attending. Prog. Brain Res. 1991;88:521–532. doi: 10.1016/s0079-6123(08)63831-5. [DOI] [PubMed] [Google Scholar]
  • 105.Foote SL, Bloom FE, Aston-Jones G. Nucleus locus coeruleus: new evidence of anatomical and physiological specificity. Physiol. Rev. 1983;63:844–914. doi: 10.1152/physrev.1983.63.3.844. [DOI] [PubMed] [Google Scholar]
  • 106.Fort P, Khateb A, Pegna A, Mϋhlethaler M, Jones BE. Noradrenergic modulation of cholinergic nucleus basalis neurons demonstrated by in vitro pharmacological and immunohistochemical evidence in the guinea-pig brain. Eur. J. Neurosci. 1995;7:1502–1511. doi: 10.1111/j.1460-9568.1995.tb01145.x. [DOI] [PubMed] [Google Scholar]
  • 107.Fritschy JM, Grzanna R. Demonstration of two separate descending noradrenergic pathways to the rat spinal cord: Evidence for an intragriseal trajectory of locus coeruleus axons in the superficial layers of the dorsal horn. J. Comp. Neurol. 1990;291:553–582. doi: 10.1002/cne.902910406. [DOI] [PubMed] [Google Scholar]
  • 108.Fritschy JM, Lyons WE, Mullen CA, Kosofsky BE, Molliver ME, Grzanna R. Brain Res. 1987;437:176–180. doi: 10.1016/0006-8993(87)91541-1. [DOI] [PubMed] [Google Scholar]
  • 109.Fruhstorfer B, Mignot E, Bowersox S, Nishino S, Dement WC, Guilleminault C. Canine narcolepsy is associated with an elevated number of alpha 2-receptors in the locus coeruleus. Brain Res. 1989;500:209–214. doi: 10.1016/0006-8993(89)90315-6. [DOI] [PubMed] [Google Scholar]
  • 110.Fu Y, Matta SG, McIntosh JM, Sharp BM. Inhibition of nicotine-induced hippocampal norepinephrine release in rats by alpha- conotoxins MII and AuIB microinjected into the locus coeruleus. Neurosci. Lett. 1999;266:113–116. doi: 10.1016/s0304-3940(99)00293-1. [DOI] [PubMed] [Google Scholar]
  • 111.Fukami H, Bradley RM. Biophysical and morphological properties of parasympathetic neurons controlling the parotid and von Ebner salivary glands in rats. J. Neurophysiol. 2005;93:678–686. doi: 10.1152/jn.00277.2004. [DOI] [PubMed] [Google Scholar]
  • 112.Fukuda A, Minami T, Nabekura J, Oomura Y. The effects of noradrenaline on neurones in the rat dorsal motor nucleus of the vagus, in vitro. J. Physiol. 1987;393:213–231. doi: 10.1113/jphysiol.1987.sp016820. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Furst S. Transmitters involved in antinociception in the spinal cord. Brain Res. Bull. 1999;48:129–141. doi: 10.1016/s0361-9230(98)00159-2. [DOI] [PubMed] [Google Scholar]
  • 114.Gallopin T, Fort P, Eggermann E, Cauli B, Luppi P-H, Rossier J, Audinat E, Mϋhlethaler M, Serafin M. Identification of sleep- promoting neurons in vitro. Nature. 2000;404:992–995. doi: 10.1038/35010109. [DOI] [PubMed] [Google Scholar]
  • 115.Gatter KC, Powell TPS. The projection of the locus coeruleus upon the neocortex in the macaque monkey. Neuroscience. 1977;2:441–445. doi: 10.1016/0306-4522(77)90009-4. [DOI] [PubMed] [Google Scholar]
  • 116.Gervasoni D, Darracq L, Fort P, Soulière F, Chouvet G, Luppi P-H. Electrophysiological evidence that noradrenergic neurons of the rat locus coeruleus are tonically inhibited by GABA during sleep. Eur. J. Neurosci. 1998;10:964–970. doi: 10.1046/j.1460-9568.1998.00106.x. [DOI] [PubMed] [Google Scholar]
  • 117.Glass MJ, Colago EEO, Pickel VM. Alpha-2A-adrenergic receptors are present on neurons in the central nucleus of the amygdala that project to the dorsal vagal complex in the rat. Synapse. 2002;46:258–268. doi: 10.1002/syn.10136. [DOI] [PubMed] [Google Scholar]
  • 118.Goddard AW, Charney DS, Germine M, Woods SW, Heninger GR, Krystal JH, Goodman WK, Price LH. Effects of tryptophan depletion on responses to yohimbine in healthy human subjects. Biol. Psychiatry. 1995;38:74–85. doi: 10.1016/0006-3223(94)00223-P. [DOI] [PubMed] [Google Scholar]
  • 119.Graham JC, Hoffman GE, Sved AF. C-Fos expression in brain in response to hypotension and hypertension in conscious rats. J. Auton. Nerv. Syst. 1995;55:92–104. doi: 10.1016/0165-1838(95)00032-s. [DOI] [PubMed] [Google Scholar]
  • 120.Grant SJ, Aston-Jones G, Redmond DE Jr. Responses of primate locus coeruleus neurons to simple and complex sensory stimuli. Brain Res. Bull. 1988;21:401–410. doi: 10.1016/0361-9230(88)90152-9. [DOI] [PubMed] [Google Scholar]
  • 121.Green GM, Lyons L, Dickenson AH. α2-adrenoceptor antagonists enhance responses of dorsal horn neurones to formalin induced inflammation. Eur. J. Pharmacol. 1998;347:201–204. doi: 10.1016/s0014-2999(98)00217-9. [DOI] [PubMed] [Google Scholar]
  • 122.Gritti I, Maiville L, Jones BE. Codistribution of GABA- with acetylcholine-synthesizing neurons in the basal forebrain of the rat. J. Comp. Neurol. 1993;329:438–457. doi: 10.1002/cne.903290403. [DOI] [PubMed] [Google Scholar]
  • 123.Grzanna R, Molliver ME. The locus coeruleus in the rat: an immunohistochemical delineation. Neuroscience. 1980;5:21–40. doi: 10.1016/0306-4522(80)90068-8. [DOI] [PubMed] [Google Scholar]
  • 124.Gurtu S, Pant KK, Sinha JN, Bhargava KP. An investigation into the mechanism of cardiovascular responses elicited by electrical stimulation of locus coeruleus and subcoeruleus in the cat. Brain Res. 1984;301:59–64. doi: 10.1016/0006-8993(84)90402-5. [DOI] [PubMed] [Google Scholar]
  • 125.Guyenet PG. Role of the ventral medulla oblongata in blood pressure regulation Central Regulation of Autonomic Functions. In: Loewy AD, Spyer KM, editors. Central Regulation of Autonomic Functions. New York: Oxford University Press; 1990. pp. 145–167. [Google Scholar]
  • 126.Guyenet PG. Central noradrenergic neurons: the autonomic connection. Prog. Brain Res. 1991;88:365–380. doi: 10.1016/s0079-6123(08)63823-6. [DOI] [PubMed] [Google Scholar]
  • 127.Guyenet PG, Stornetta RL, Riley T, Norton FR, Rosin DL, Lynch KR. Alpha 2A-adrenergic receptors are present in lower brainstem catecholaminergic and serotonergic neurons innervating spinal cord. Brain Res. 1994;638:285–294. doi: 10.1016/0006-8993(94)90661-0. [DOI] [PubMed] [Google Scholar]
  • 128.Guyenet PG, Young BS. Projections of nucleus paragigantocellularis lateralis to locus coeruleus and other structures in rat. Brain Res. 1987;406:171–184. doi: 10.1016/0006-8993(87)90781-5. [DOI] [PubMed] [Google Scholar]
  • 129.Haas H, Panula P. The role of histamine and the tuberomamillary nucleus in the nervous system. Nat. Rev. Neurosci. 2003;4:121–130. doi: 10.1038/nrn1034. [DOI] [PubMed] [Google Scholar]
  • 130.Hagan JJ, Leslie RA, Patel S, Evans ML, Wattam TA, Holmes S, Benham CD, Taylor SG, Routledge C, Hemmati P, Munton RP, Ashmeade TE, Shah AS, Hatcher JP, Hatcher PD, Jones DNC, Smith MI, Piper DC, Hunter AJ, Porter RA, Upton N. Orexin A activates locus coeruleus cell firing and increases arousal in the rat. Proc. Natl. Acad. Sci. USA. 1999;96:10911–10916. doi: 10.1073/pnas.96.19.10911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Hancock MB, Fougerousse CL. Spinal projections from the nucleus locus coeruleus and nucleus subcoeruleus in the cat and monkey as demonstrated by the retrograde transport of horseradish peroxidase. Brain Res. Bull. 1976;1:229–234. doi: 10.1016/0361-9230(76)90072-1. [DOI] [PubMed] [Google Scholar]
  • 132.Head GA, Chan CKS, Burke SL. Relationship between imidazoline and α 2-adrenoceptors involved in the sympatho-inhibitory actions of centrally acting antihypertensive agents. J. Auton. Nerv. Syst. 1998;72:163–169. doi: 10.1016/s0165-1838(98)00101-5. [DOI] [PubMed] [Google Scholar]
  • 133.Heal DJ, Cheetham SC, Butler SA, Gosden J, Prow MR, Buckett WR. Receptor binding and functional evidence suggest that postsynaptic alpha 2-adrenoceptors in rat brain are of the alpha 2D subtype. Eur. J. Pharmacol. 1995;277:215–221. doi: 10.1016/0014-2999(95)00078-y. [DOI] [PubMed] [Google Scholar]
  • 134.Heal DJ, Prow MR, Butler SA, Buckett WR. Mediation of mydriasis in conscious rats by central postsynaptic α2-adrenoceptors. Pharmacol. Biochem. Behav. 1995;50:219–224. doi: 10.1016/0091-3057(94)00299-x. [DOI] [PubMed] [Google Scholar]
  • 135.Hentall ID, Pinzon A, Noga BR. Spatial and temporal patterns of serotonin release in the rat’s lumbar spinal cord following electrical stimulation of the nucleus raphe magnus. Neuroscience. 2006;142:893–903. doi: 10.1016/j.neuroscience.2006.06.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Hermann DM, Luppi P-H, Peyron C, Hinckel P, Jouvet M. Afferent projections to the rat nuclei raphe magnus, raphe pallidus and reticularis gigantocellularis pars α demonstrated by iontophoretic application of choleratoxin (subunit b) J. Chem. Neuroanat. 1997;13:1–21. doi: 10.1016/s0891-0618(97)00019-7. [DOI] [PubMed] [Google Scholar]
  • 137.Holstege G. Some anatomical observations on the projections from the hypothalamus to brainstem and spinal cord: an HRP and autoradiographic tracing study in the cat. J. Comp. Neurol. 1987;260:98–126. doi: 10.1002/cne.902600109. [DOI] [PubMed] [Google Scholar]
  • 138.Horvath TL, Peyron C, Diano S, Ivanov A, Aston-Jones G, Kilduff TS, van den Pol AN. Hypocretin (orexin) activation and synaptic innervation of the locus coeruleus noradrenergic system. J. Comp. Neurol. 1999;415:145–159. [PubMed] [Google Scholar]
  • 139.Hou RH, Freeman C, Langley RW, Szabadi E, Bradshaw CM. Does modafinil activate the locus coeruleus in man? Comparison of modafinil and clonidine on arousal and autonomic functions in human volunteers. Psychopharmacology (Berl.) 2005;181:537–549. doi: 10.1007/s00213-005-0013-8. [DOI] [PubMed] [Google Scholar]
  • 140.Hou YP, Manns ID, Jones BE. Immunostaining of cholinergic pontomesencephalic neurons for α1 versus α2 adrenergic receptors suggests different sleep-wake state activities and roles. Neuroscience. 2002;114:517–521. doi: 10.1016/s0306-4522(02)00340-8. [DOI] [PubMed] [Google Scholar]
  • 141.Hrabovszky E, Liposits Z. Adrenergic innervation of dopamine neurons in the hypothalamic arcuate nucleus of the rat. Neurosci. Lett. 1994;182:143–146. doi: 10.1016/0304-3940(94)90783-8. [DOI] [PubMed] [Google Scholar]
  • 142.Huang H-P, Wang S-R, Yao W, Zhang C, Zhou Y, Chen X-W, Zhang B, Xiong W, Wang L-Y, Zheng L-H. Landry, M. Hkfelt, T. Xu, Z-Q D, Zhou Z. Long latency of evoked quantal transmitter release from somata of locus coeruleus neurons in rat pontine slices. Proc. Natl. Acad. Sci. USA. 2007;104:1401–1406. doi: 10.1073/pnas.0608897104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Hudson AL, Mallard NJ, Tyacke R, Nutt DJ. [3H]-RX821002: A highly selective ligand for the identification of alpha2-adrenoreceptors in the rat brain. Mol. Neuropharmacol. 1992;1:219–229. [Google Scholar]
  • 144.Hume SP, Lammertsma AA, Opacka-Juffry J, Ahier RG, Myers R, Cremer JE, Hudson AL, Nutt DJ, Pike VW. Quantification of in vivo binding of [3H]-RX821002 in rat brain: Evaluation as a radioligand for central alpha2-adrenoceptors. Nucl. Med. Biol. 1992;19:841–849. doi: 10.1016/0883-2897(92)90170-4. [DOI] [PubMed] [Google Scholar]
  • 145.Hwang K-R, Chan SHH, Chan JYH. Noradrenergic neurotransmission at PVN in locus coeruleus-induced baroreflex suppression in rats. Heart Circ. Physiol. 1998;43:H1284–1292. doi: 10.1152/ajpheart.1998.274.4.H1284. [DOI] [PubMed] [Google Scholar]
  • 146.Illes P, Norenberg W. Blockade of alpha 2-adrenoceptors increases opioid mu-receptor-mediated inhibition of the firing rate of rat locus coeruleus neurones. Naunyn Schmiedebergs Arch. Pharmacol. 1990;342:490–496. doi: 10.1007/BF00169034. [DOI] [PubMed] [Google Scholar]
  • 147.Ishikawa M, Tanaka C. Morphological organization of catecholamine terminals in the diencephalons of the rhesus monkey. Brain Res. 1977;119:43–55. doi: 10.1016/0006-8993(77)90090-7. [DOI] [PubMed] [Google Scholar]
  • 148.Ivanov A, Aston-Jones G. Extranuclear dendrites of locus coeruleus neurons: activation by glutamate and modulation of activity by alpha adrenoceptors. J. Neurophysiol. 1995;74:2427–2436. doi: 10.1152/jn.1995.74.6.2427. [DOI] [PubMed] [Google Scholar]
  • 149.Iwase M, Homma I, Shioda S, Nakai Y. Histamine immunoreactive neurons in the brain stem of the rabbit. Brain Res. Bull. 1993;32:267–272. doi: 10.1016/0361-9230(93)90187-g. [DOI] [PubMed] [Google Scholar]
  • 150.Jacobs BL, Martin-Cora FJ, Fornal CA. Activity of medullary serotonergic neurons in freely moving animals. Brain Res. Rev. 2002;40:45–52. doi: 10.1016/s0165-0173(02)00187-x. [DOI] [PubMed] [Google Scholar]
  • 151.Jansen AS, Ter Horst GJ, Mettenleiter TC, Loewy AD. CNS cell groups projecting to the submandibular parasympathetic preganglionic neurons in the rat: a retrograde transneuronal viral cell body labeling study. Brain Res. 1992;572:253–260. doi: 10.1016/0006-8993(92)90479-s. [DOI] [PubMed] [Google Scholar]
  • 152.Jasmin L, Tien D, Weinshenker D, Palmiter RD, Green PG, Janni G, O’Hara PT. The NK1 receptor mediates both the hyperalgesia and the resistance to morphine in mice lacking noradrenaline. Proc. Natl. Acad. Sci. USA. 2002;99:1029–1034. doi: 10.1073/pnas.012598599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Jodo E, Aston-Jones G. Activation of locus coeruleus by prefrontal cortex is mediated by excitatory amino acid inputs. Brain Res. 1997;768:327–332. doi: 10.1016/s0006-8993(97)00703-8. [DOI] [PubMed] [Google Scholar]
  • 154.Jodo E, Chiang C, Aston-Jones G. Potent excitatory influence of prefrontal cortex activity on noradrenergic locus coeruleus neurons. Neuroscience. 1998;83:3–79. doi: 10.1016/s0306-4522(97)00372-2. [DOI] [PubMed] [Google Scholar]
  • 155.Johnson AD, Peoples J, Stornetta RL, Van Bockstaele EJ. Opioid circuits originating from the nucleus paragigantocellularis and their potential role in opiate withdrawal. Brain Res. 2002;955:72–84. doi: 10.1016/s0006-8993(02)03367-x. [DOI] [PubMed] [Google Scholar]
  • 156.Jones BE. Immunohistochemical study of choline acetyltransferase- immunoreactive processes and cells innervating the pontomedullary reticular formation in the rat. J. Comp. Neurol. 1990;295:485–514. doi: 10.1002/cne.902950311. [DOI] [PubMed] [Google Scholar]
  • 157.Jones BE. Paradoxical sleep and its chemical/structural substrates in the brain. Neuroscience. 1991;40:637–656. doi: 10.1016/0306-4522(91)90002-6. [DOI] [PubMed] [Google Scholar]
  • 158.Jones BE. Basic mechanisms of sleep-wake state. In: Kryger MH, Roth T, Dement WC, editors. Principles and Practice of Sleep Medicine. 3rd Ed. Philadelphia: W.B. Saunders Company; 2000. pp. 134–154. [Google Scholar]
  • 159.Jones BE. Activity, modulation and role of basal forebrain cholinergic neurons innervating the cerebral cortex. Prog. Brain Res. 2004;145:157–169. doi: 10.1016/S0079-6123(03)45011-5. [DOI] [PubMed] [Google Scholar]
  • 160.Jones BE. From waking to sleeping: neuronal and chemical substrates. Trends Pharmacol. Sci. 2005;26:578–586. doi: 10.1016/j.tips.2005.09.009. [DOI] [PubMed] [Google Scholar]
  • 161.Jones BE, Harper ST, Halaris AE. Effects of locus coeruleus lesions upon cerebral monoamine content, sleep-wakefulness states and the response to amphetamine in the cat. Brain Res. 1977;124:473–496. doi: 10.1016/0006-8993(77)90948-9. [DOI] [PubMed] [Google Scholar]
  • 162.Jones SL, Light AR. Serotoninergic medullary raphe spinal projection to the lumbar spinal cord in the rat: a retrograde immunohistochemical study. J. Comp. Neurol. 1992;322:599–610. doi: 10.1002/cne.903220413. [DOI] [PubMed] [Google Scholar]
  • 163.Jones BE, Moore RY. Ascending projections of the locus coeruleus in the rat. II Autoradiographic study. Brain Res. 1977;127:25–53. [PubMed] [Google Scholar]
  • 164.Jones BE, Yang T-Z. The efferent projections from the reticular formation and the locus coeruleus studies by anterograde and retrograde axonal transport in the rat. J. Comp. Neurol. 1985;242:56–92. doi: 10.1002/cne.902420105. [DOI] [PubMed] [Google Scholar]
  • 165.Kaitin KI, Bliwise DL, Gleason C, Nino-Murcia G, Dement WC, Libet B. Sleep disturbance produced by electrical stimulation of the locus coeruleus in a human subject. Biol. Psychiatry. 1986;21:710–716. doi: 10.1016/0006-3223(86)90235-0. [DOI] [PubMed] [Google Scholar]
  • 166.Kalia M. Neurobiology of sleep. Metab. Clin. Exp. 2006;55 (suppl. 2):S2– S6. doi: 10.1016/j.metabol.2006.07.005. [DOI] [PubMed] [Google Scholar]
  • 167.Kang Y-M, Ouyang W, Chen J-Y, Qiao J-T, Dafny N. Norepinephrine modulates single hypothalamic arcuate neurons via α1 and β adrenergic receptors. Brain Res. 2000;869:146–157. doi: 10.1016/s0006-8993(00)02380-5. [DOI] [PubMed] [Google Scholar]
  • 168.Kapoor V, Minson J, Chalmers J. Ventral medulla stimulation increases blood pressure and spinal cord amino acid release. Neuroreport. 1992;3:55–58. doi: 10.1097/00001756-199201000-00014. [DOI] [PubMed] [Google Scholar]
  • 169.Kaur S, Saxena RN, Mallick BN. GABAergic neurons in prepositus hypoglossi regulate REM sleep by its action on locus coeruleus in freely moving rats. Synapse. 2001;42:141–150. doi: 10.1002/syn.1109. [DOI] [PubMed] [Google Scholar]
  • 170.Kawahara Y, Kawahara H, Westerink BHC. Tonic regulation of the activity of noradrenergic neurons in the locus coeruleus of the conscious rat studied by dual-probe microdialysis. Brain Res. 1999;823:42–48. doi: 10.1016/s0006-8993(99)01062-8. [DOI] [PubMed] [Google Scholar]
  • 171.Kayama Y, Ohta M, Jodo E. Firing of “possibly” cholinergic neurons in the rat laterodorsal tegmental nucleus during sleep and wakefulness. Brain Res. 1992;569:210–220. doi: 10.1016/0006-8993(92)90632-j. [DOI] [PubMed] [Google Scholar]
  • 172.Key BJ, Krzywoskinski L. Electrocortical changes induced by the perfusion of noradrenaline, acetylcholine and their antagonists directly into the dorsal raphe nucleus of the cat. Br. J. Pharmacol. 1977;61:297–305. doi: 10.1111/j.1476-5381.1977.tb08419.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Khokhlova ON, Murashev AN, Medvedev OS. Role of the I1- imidazoline receptors and α2-adrenoceptors in hemodynamic effects of moxonidine administration into the rostroventrolateral medulla. Bull. Exp. Biol. Med. 2001;131:336–339. doi: 10.1023/a:1017943901070. [DOI] [PubMed] [Google Scholar]
  • 174.Kiernan JA. Maryland, USA: Lippincott Williams & Wilkins; 2005. Barr’s the human nervous system: an anatomical viewpoint. [Google Scholar]
  • 175.Kim M, Chiego DJ, Bradley RM. Morphology of parasympathetic neurons innervating rat lingual salivary glands. Auton. Neurosci. 2004;111:27–36. doi: 10.1016/j.autneu.2004.01.006. [DOI] [PubMed] [Google Scholar]
  • 176.Kim M-A, Lee HS, Lee BY, Waterhouse BD. Reciprocal connections between subdivisions of the dorsal raphe and the nuclear core of the locus coeruleus in the rat. Brain Res. 2004;1026:56–67. doi: 10.1016/j.brainres.2004.08.022. [DOI] [PubMed] [Google Scholar]
  • 177.Kitahama K, Nagatsu I, Geffard M, Maeda T. Distribution of dopamine-immunoreactive fibers in the rat brainstem. J. Chem. Neuroanat. 2000;18:1–9. doi: 10.1016/s0891-0618(99)00047-2. [DOI] [PubMed] [Google Scholar]
  • 178.Kiyashchenko LI, Mileykovskiy BY, Lai Y-Y, Siegel JM. Increased and decreased muscle tone with orexin (hypocretin) microinjections in the locus coeruleus and pontine inhibitory area. J. Neurophysiol. 2001;85:2008– 2016. doi: 10.1152/jn.2001.85.5.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Kiyokawa Y, Kikusui T, Takeuchi Y, Mori Y. Mapping the neural circuit activated by alarm pheromone perception by c-Fos immunohistochemistry. Brain Res. 2005;1043:145–154. doi: 10.1016/j.brainres.2005.02.061. [DOI] [PubMed] [Google Scholar]
  • 180.Klepper A, Herbert H. Distribution and origin of noradrenergic and serotonergic fibers in the cochlear nucleus and inferior colliculus of the rat. Brain Res. 1991;557:190–201. doi: 10.1016/0006-8993(91)90134-h. [DOI] [PubMed] [Google Scholar]
  • 181.Klooster J, Vrensen GF, Muller LJ, van der Want JJ. Efferent projections of the olivary pretectal nucleus in the albino rat subserving the pupillary light reflex and related reflexes. A light microscopic tracing study. Brain Res. 1995;688:34–46. doi: 10.1016/0006-8993(95)00497-e. [DOI] [PubMed] [Google Scholar]
  • 182.Ko EM, Estabrooke IV, McCarthy M, Scammell TE. Wake- related activity of Tuberomammillary neurons in rats. Brain Res. 2003;992:220– 226. doi: 10.1016/j.brainres.2003.08.044. [DOI] [PubMed] [Google Scholar]
  • 183.Korf J, Bunney BS, Aghajanian GK. Noradrenergic neurons: morphine inhibition of spontaneous activity. Eur. J. Pharmacol. 1974;25:165–169. doi: 10.1016/0014-2999(74)90045-4. [DOI] [PubMed] [Google Scholar]
  • 184.Krauchi K, Wirz-Justice A, Morimasa T, Willener R, Feer H. Hypothalamic alpha 2- and beta-adrenoceptor rhythms are correlated with circadian feeding: evidence from chronic methamphetamine treatment and withdrawal. Brain Res. 1984;321:83–90. doi: 10.1016/0006-8993(84)90683-8. [DOI] [PubMed] [Google Scholar]
  • 185.Kromer LF, Moore RY. Cochlear nucleus innervation by central norepinephrine neurons in the rat. Brain Res. 1976;118:531–537. doi: 10.1016/0006-8993(76)90327-9. [DOI] [PubMed] [Google Scholar]
  • 186.Kromer LF, Moore RY. Norepinephrine innervation of the cochlear nuclei by locus coeruleus neurons in the rat. Anat. Embryol. (Berl.) 1980;158:227–244. doi: 10.1007/BF00315908. [DOI] [PubMed] [Google Scholar]
  • 187.Kumari V, Cotter P, Corr PJ, Gray JA, Checkley SA. Effect of clonidine on the human acoustic startle reflex. Psychopharmacology (Berl.) 1996;123:353–360. doi: 10.1007/BF02246646. [DOI] [PubMed] [Google Scholar]
  • 188.Kwiat GC, Basbaum AI. The origin of brainstem noradrenergic and serotonergic projections to the spinal cord dorsal horn in the rat. Somatosens. Mot. Res. 1992;9:157–173. doi: 10.3109/08990229209144768. [DOI] [PubMed] [Google Scholar]
  • 189.Lang B, Li H, Kang J-F, Li Y-Q. Alpha-2 adrenoceptor mediating the facilitatory effect of norepinephrine on the glycine response in the spinal dorsal horn neuron of the rat. Life Sci. 2003;73:893–905. doi: 10.1016/s0024-3205(03)00352-7. [DOI] [PubMed] [Google Scholar]
  • 190.Langer T, Fuchs AF, Chubb MC, Scudder CA, Lisberger SG. Floccular efferents in the rhesus macaque as revealed by autoradiography and horseradish peroxidase. J. Comp. Neurol. 1985;235:26–37. doi: 10.1002/cne.902350103. [DOI] [PubMed] [Google Scholar]
  • 191.Lapiz MDS, Morilak DA. Noradrenergic modulation of cognitive function in rat medial prefrontal cortex as measured by attentional set shifting capability. Neuroscience. 2006;137:1039–1049. doi: 10.1016/j.neuroscience.2005.09.031. [DOI] [PubMed] [Google Scholar]
  • 192.LeDoux J. Fear and the brain: where have we been, and where are we going? Biol. Psychiatry. 1998;44:1229–1238. doi: 10.1016/s0006-3223(98)00282-0. [DOI] [PubMed] [Google Scholar]
  • 193.Lee HS, Kim M-A, Waterhouse BD. Retrograde double-labeling study of common afferent projections to the dorsal raphe and the nuclear core of the locus coeruleus in the rat. J. Comp. Neurol. 2005;481:179–193. doi: 10.1002/cne.20365. [DOI] [PubMed] [Google Scholar]
  • 194.Lee HS, Lee BY, Waterhouse BD. Retrograde study of projections from the tuberomammillary nucleus to the dorsal raphe and the locus coeruleus in the rat. Brain Res. 2005;1043:65–75. doi: 10.1016/j.brainres.2005.02.050. [DOI] [PubMed] [Google Scholar]
  • 195.Lee HS, Park SH, Song WC, Waterhouse BD. Retrograde study of hypocretin-1 (orexin-A) projections to subdivisions of the dorsal raphe nucleus in the rat. Brain Res. 2005;1059:35–45. doi: 10.1016/j.brainres.2005.08.016. [DOI] [PubMed] [Google Scholar]
  • 196.Leong SK, Shieh JY, Wong WC. Localizing spinalcord- projecting neurons in adult albino rats. J. Comp. Neurol. 1984;228:1–17. doi: 10.1002/cne.902280103. [DOI] [PubMed] [Google Scholar]
  • 197.Levitt P, Moore RY. Origin and organization of brainstem catecholamine innervation in the rat. J. Comp. Neurol. 1979;186:505–528. doi: 10.1002/cne.901860402. [DOI] [PubMed] [Google Scholar]
  • 198.Levitt P, Rakic P, Goldman-Rakic P. Region-specific distribution of catecholamine afferents in primate cerebral cortex: a fluorescence histochemical analysis. J. Comp. Neurol. 1984;227:23–36. doi: 10.1002/cne.902270105. [DOI] [PubMed] [Google Scholar]
  • 199.Lewis DI, Coote JH. Excitation and inhibition of rat sympathetic preganglionic neurones by catecholamines. Brain Res. 1990;530:229–234. doi: 10.1016/0006-8993(90)91287-q. [DOI] [PubMed] [Google Scholar]
  • 200.Li D-P, Atnip LM, Chen S-R, Pan H-L. Regulation of synaptic inputs to paraventricular-spinal output neurons by α2 adrenergic receptors. J. Neurophysiol. 2005;93:393–402. doi: 10.1152/jn.00564.2004. [DOI] [PubMed] [Google Scholar]
  • 201.Li X, Rainnie DG, McCarley RW, Greene RW. Presynaptic nicotinic receptors facilitate monoaminergic transmission. J. Neurosci. 1998;18:1904–1912. doi: 10.1523/JNEUROSCI.18-05-01904.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Li Y, van den Pol AN. Direct and indirect inhibition by catecholamines of hypocretin/orexin neurons. J. Neurosci. 2005;25:173–183. doi: 10.1523/JNEUROSCI.4015-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Lightman SL, Todd K, Everitt BJ. Ascending noradrenergic projections from the brainstem: evidence for a major role in the regulation of blood pressure and vasopressin secretion. Exp. Brain Res. 1984;55:145–151. doi: 10.1007/BF00240508. [DOI] [PubMed] [Google Scholar]
  • 204.Lindvall O, Stenevi U. Dopamine and noradrenaline neurons projecting to the septal area in the rat. Cell Tissue Res. 1978;190:383–407. doi: 10.1007/BF00219554. [DOI] [PubMed] [Google Scholar]
  • 205.Loewenfeld IE. The Pupil: Anatomy, Physiology, and Clinical Applications. Detroit, Michigan: Wayne State University Press; 1993. [Google Scholar]
  • 206.Loewy AD. Raphe pallidus and raphe obscurus projections to the intermediolateral cell column in the rat. Brain Res. 1981;222:129–133. doi: 10.1016/0006-8993(81)90946-x. [DOI] [PubMed] [Google Scholar]
  • 207.Loewy AD, Araujo JC, Kerr FWL. Pupillodilator pathways in the brain stem of the cat: anatomical and electrophysiological identification of a central autonomic pathway. Brain Res. 1973;60:65–91. doi: 10.1016/0006-8993(73)90851-2. [DOI] [PubMed] [Google Scholar]
  • 208.Loughlin SE, Foote SL, Bloom FE. Efferent projections of nucleus locus coeruleus: topographic organization of cells of origin demonstrated by three-dimensional reconstruction. Neuroscience. 1986;18:291–306. doi: 10.1016/0306-4522(86)90155-7. [DOI] [PubMed] [Google Scholar]
  • 209.Loughlin SE, Foote SL, Fallon JH. Locus coeruleus projections to cortex: topography, morphology, and collateralization. Brain Res. Bull. 1982;9:287–294. doi: 10.1016/0361-9230(82)90142-3. [DOI] [PubMed] [Google Scholar]
  • 210.Loughlin SE, Foote SL, Grzanna R. Efferent projections of nucleus locus coeruleus: morphologic subpopulations have different efferent targets. Neuroscience. 1986;18:307–319. doi: 10.1016/0306-4522(86)90156-9. [DOI] [PubMed] [Google Scholar]
  • 211.Lu J, Jhou TC, Saper CB. Identification of wake-active dopaminergic neurons in the ventral periaqueductal gray matter. J. Neurosci. 2006;26:193–202. doi: 10.1523/JNEUROSCI.2244-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Lu J, Sherman D, Devor M, Saper CB. A putative flip-flop switch for control of REM sleep. Nature. 2006;441:589–594. doi: 10.1038/nature04767. [DOI] [PubMed] [Google Scholar]
  • 213.Luiten PGM, Ter Horst GJ, Karst H, Steffens AB. The course of the paraventricular hypothalamic efferents to autonomic structures in medulla and spinal cord. Brain Res. 1985;329:374–378. doi: 10.1016/0006-8993(85)90554-2. [DOI] [PubMed] [Google Scholar]
  • 214.Lung MA. Mechanisms of sympathetic enhancement and inhibition of parasympathetically induced salivary secretion in anaesthetized dogs. Br. J. Pharmacol. 1994;112:411–416. doi: 10.1111/j.1476-5381.1994.tb13087.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Luppi P-H, Aston-Jones G, Akaoka H, Chouvet G, Jouvet M. Afferent projections to the rat locus coeruleus demonstrated by retrograde and anterograde tracing with cholera-toxin B subunit and Phaseolus vulgaris leucoagglutinin. Neuroscience. 1995;65:119–160. doi: 10.1016/0306-4522(94)00481-j. [DOI] [PubMed] [Google Scholar]
  • 216.Luppi P-H, Gervasoni D, Peyron C, Rampon C, Barbagli B, Boissard R, Fort P. Norepinephrine and REM sleep Rapid Eye Movement Sleep. Rapid Eye Movement Sleep. In: Mallick BN, Inoué S, editors. New Delhi India: Narosa Publishing House; 1999. pp. 107–122. [Google Scholar]
  • 217.Lyons WE, Grzanna R. Noradrenergic neurons with divergent projections to the motor trigeminal nucleus and the spinal cord: a double retrograde neuronal labeling study. Neuroscience. 1988;26:681–693. doi: 10.1016/0306-4522(88)90174-1. [DOI] [PubMed] [Google Scholar]
  • 218.Machado BH, Brody MJ. Mechanisms of pressor response produced by stimulation of nucleus ambiguus. Am. J. Physiol. 1990;259:R955–R962. doi: 10.1152/ajpregu.1990.259.5.R955. [DOI] [PubMed] [Google Scholar]
  • 219.Maeda T, Kitahama K, Geffard M. Dopaminergic innervation of rat locus coeruleus: a light and electron microscopic immunohistochemical study. Microsc. Res. Tech. 1994;29:211–218. doi: 10.1002/jemt.1070290306. [DOI] [PubMed] [Google Scholar]
  • 220.Mair RD, Zhang Y, Bailey KR, Toupin MM, Mair RG. Effects of clonidine in the locus coeruleus on prefrontal- and hippocampal-dependent measures of attention and memory in the rat. Psychopharmacology (Berl.) 2005;181:280–288. doi: 10.1007/s00213-005-2263-x. [DOI] [PubMed] [Google Scholar]
  • 221.Manns ID, Alonso A, Jones BE. Discharge profiles of juxtacellularly labelled and immunohistochemically identified GABAergic basal forebrain neurons recorded in association with the electroencephalogram in anesthetized rats. J. Neurosci. 2000;20:9252–9263. doi: 10.1523/JNEUROSCI.20-24-09252.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Manns ID, Alonso A, Jones BE. Discharge properties of juxtacellularly labelled and immunohistochemically identified cholinergic basal forebrain neurons recorded in association with the electroencephalogram in anesthetized rats. J. Neurosci. 2000;20:1505–1518. doi: 10.1523/JNEUROSCI.20-04-01505.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Manns ID, Lee MG, Modirrousta M, Hou YP, Jones BE. Alpha 2 adrenergic receptors on GABAergic, putative sleep-promoting basal forebrain neurons. Eur. J. Neurosci. 2003;18:723–727. doi: 10.1046/j.1460-9568.2003.02788.x. [DOI] [PubMed] [Google Scholar]
  • 224.Marchand ER, Riley JN, Moore RY. Interpeduncular nucleus afferents in the rat. Brain Res. 1980;193:339–352. doi: 10.1016/0006-8993(80)90169-9. [DOI] [PubMed] [Google Scholar]
  • 225.Marchenko V, Sapru HN. Cardiovascular responses to chemical stimulation of the lateral tegmental field and adjacent medullary reticular formation in the rat. Brain Res. 2003;977:247–260. doi: 10.1016/s0006-8993(03)02719-7. [DOI] [PubMed] [Google Scholar]
  • 226.Martinez-Peña y Valenzuela I, Rogers RC, Hermann GE, Travagli RA. Norepinephrine effects on identified neurons of the rat dorsal motor nucleus of the vagus. Am. J. Physiol. Gastrointest. Liver Physiol. 2004;286:G333– G339. doi: 10.1152/ajpgi.00289.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Mathis J, Hess CW, Bassetti C. Isolated mediotegmental lesion causing narcolepsy and rapid eye movement sleep behaviour disorder: a case evidencing a common pathway in narcolepsy and rapid eye movement sleep behaviour disorder. J. Neurol. Neurosurg. Psychiatry. 2007;78:427–429. doi: 10.1136/jnnp.2006.099515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Matsuo S-I, Jang I-S, Nabekura J, Akaike N. α2-adrenoceptor- mediated presynaptic modulation of GABAergic transmission in mechanically dissociated rat ventrolateral preoptic neurons. J. Neurophysiol. 2003;89:1640– 1648. doi: 10.1152/jn.00491.2002. [DOI] [PubMed] [Google Scholar]
  • 229.Matsuo R, Morimoto T, Kang Y. Neural activity of the superior salivatory nucleus in rats. Eur. J. Morphol. 1998;36(Suppl.):203–207. [PubMed] [Google Scholar]
  • 230.Matsutani K, Tsuruoka M, Shinya A, Furuya R, Kawawa T. Stimulation of the locus coeruleus suppresses trigeminal sensorimotor function in the rat. Brain Res. Bull. 2000;53:827–832. doi: 10.1016/s0361-9230(00)00426-3. [DOI] [PubMed] [Google Scholar]
  • 231.McBride RL, Sutin J. Projections of the locus coeruleus and adjacent pontine tegmentum in the cat. J. Comp. Neurol. 1976;165:265–284. doi: 10.1002/cne.901650302. [DOI] [PubMed] [Google Scholar]
  • 232.McCormick DA, Bal T. Sleep and arousal: thalamocortical mechanisms. Ann. Rev. Neurosci. 1997;20:185–215. doi: 10.1146/annurev.neuro.20.1.185. [DOI] [PubMed] [Google Scholar]
  • 233.McCormick DA, Pape HC, Williamson A. Actions of norepinephrine in the cerebral cortex and thalamus: implications for function of the central noradrenergic system. Prog. Brain Res. 1991;88:293–305. doi: 10.1016/s0079-6123(08)63817-0. [DOI] [PubMed] [Google Scholar]
  • 234.McDougle CJ, Krystal JH, Price LH, Heninger GR, Charney DS. Noradrenergic response to acute ethanol administration in healthy subjects: comparison with intravenous yohimbine. Psychopharmacology (Berl.) 1995;118:127–135. doi: 10.1007/BF02245830. [DOI] [PubMed] [Google Scholar]
  • 235.McGinty DJ, Harper RM. Dorsal raphe neurons: depression of firing during sleep in cats. Brain Res. 1976;101:569–575. doi: 10.1016/0006-8993(76)90480-7. [DOI] [PubMed] [Google Scholar]
  • 236.McKitrick DJ, Calaresu FR. Nucleus ambiguus inhibits activity of cardiovascular units in RVLM. Brain Res. 1996;742:203–210. doi: 10.1016/s0006-8993(96)00971-7. [DOI] [PubMed] [Google Scholar]
  • 237.McRae-Degueurce A, Milon H. Serotonin and dopamine afferents to the rat locus coeruleus: a biochemical study after lesioning of the ventral mesencephalic tegmental-A10 region and the raphe dorsalis. Brain Res. 1983;263:344–347. doi: 10.1016/0006-8993(83)90327-x. [DOI] [PubMed] [Google Scholar]
  • 238.Mieda M, Yanagisawa M. Sleep, feeding, and neuropeptides: roles of orexins and orexin receptors. Curr. Opin. Neurobiol. 2002;12:339–345. doi: 10.1016/s0959-4388(02)00331-8. [DOI] [PubMed] [Google Scholar]
  • 239.Millan MJ. Descending control of pain. Prog. Neurobiol. 2002;66:355–474. doi: 10.1016/s0301-0082(02)00009-6. [DOI] [PubMed] [Google Scholar]
  • 240.Milon H, McRae-Degueurce A. Pharmacological investigation on the role of dopamine in the rat locus coeruleus. Neurosci. Lett. 1982;30:297–301. doi: 10.1016/0304-3940(82)90416-5. [DOI] [PubMed] [Google Scholar]
  • 241.Mizukami T. Immunocytochemical localization of beta2-adrenergic receptors in the rat spinal cord and their spatial relationships to tyrosine hydroxylase-immunoreactive terminals. Kurume Med. J. 2004;51:175–183. doi: 10.2739/kurumemedj.51.175. [DOI] [PubMed] [Google Scholar]
  • 242.Modirrousta M, Mainville L, Jones BE. GABAergic neurons with α2-adrenergic receptors in basal forebrain and preoptic area express c-Fos during sleep. Neuroscience. 2004;129:803–810. doi: 10.1016/j.neuroscience.2004.07.028. [DOI] [PubMed] [Google Scholar]
  • 243.Moore RY, Bloom FE. Central catecholamine neuron systems: anatomy and physiology of the norepinephrine and epinephrine systems. Ann. Rev. Neurosci. 1979;2:113–168. doi: 10.1146/annurev.ne.02.030179.000553. [DOI] [PubMed] [Google Scholar]
  • 244.Morgan CA, Southwick SM, Grillon C, Davis M, Krystal JH, Charney DS. Yohimbine-facilitated acoustic startle reflex in humans. Psychopharmacology (Berl.) 1993;110:342–346. doi: 10.1007/BF02251291. [DOI] [PubMed] [Google Scholar]
  • 245.Morrison SF. RVLM and raphe differentially regulate sympathetic outflows to splanchnic and brown adipose tissue. Am. J. Physiol. 1999;276:R962– R973. doi: 10.1152/ajpregu.1999.276.4.R962. [DOI] [PubMed] [Google Scholar]
  • 246.Morrison JH, Foote SL, O’Connor D, Bloom FE. Laminar, tangential and regional organization of the noradrenergic innervation of monkey cortex: dopamine-β-hydroxylase immunohistochemistry. Brain Res. Bull. 1982;9:309–319. doi: 10.1016/0361-9230(82)90144-7. [DOI] [PubMed] [Google Scholar]
  • 247.Morrison JH, Molliver ME, Grzanna R, Coyle JT. Noradrenergic innervation patterns in three regions of medial cortex: an immunofluorescence characterization. Brain Res. Bull. 1979;4:849–857. doi: 10.1016/0361-9230(79)90022-4. [DOI] [PubMed] [Google Scholar]
  • 248.Moruzzi G, Magoun HW. Brain stem reticular formation and activation of the EEG. Electroenceph. Clin. Neurophysiol. 1949;1:455–473. [PubMed] [Google Scholar]
  • 249.Müller EE, Locatelli V, Cocchi D. Neuroendocrine control of growth hormone secretion. Physiol. Rev. 1999;79:511–607. doi: 10.1152/physrev.1999.79.2.511. [DOI] [PubMed] [Google Scholar]
  • 250.Murakami T, Morita Y. Morphology and distribution of the projection neurons in the cerebellum in a teleost, Sebastiscus marmoratus. J. Comp. Neurol. 1987;256:607–623. doi: 10.1002/cne.902560413. [DOI] [PubMed] [Google Scholar]
  • 251.Murase S, Inui K, Nosaka S. Baroreceptor inhibition of the locus coeruleus noradrenergic neurons. Neuroscience. 1994;61:635–643. doi: 10.1016/0306-4522(94)90440-5. [DOI] [PubMed] [Google Scholar]
  • 252.Myers EA, Banihashemi L, Rinaman L. The anxiogenic drug yohimbine activates central viscerosensory circuits in rats. J. Comp. Neurol. 2005;492:426–441. doi: 10.1002/cne.20727. [DOI] [PubMed] [Google Scholar]
  • 253.Nagao S, Kitamura T, Nakamura N, Hiramatsu T, Yamada J. Location of efferent terminals of the primate flocculus and ventral paraflocculus revealed by anterograde axonal transport methods. Neurosci. Res. 1997;27:257–269. doi: 10.1016/s0168-0102(97)01160-7. [DOI] [PubMed] [Google Scholar]
  • 254.Nakamura K, Matsumura K, Hübschle T, Nakamura Y, Hioki H, Fujiyama F, Boldogköi Z, König M, Thiel H-J, Gerstberger R, Kobayashi S, Kaneko T. Identification of sympathetic premotor neurons in medullary raphe regions mediating fever and other thermoregulatory functions. J. Neurosci. 2004;24:5370–5380. doi: 10.1523/JNEUROSCI.1219-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Nakamura K, Matsumura K, Kobayashi S, Kaneko T. Sympathetic premotor neurons mediating thermoregulatory function. Neurosci. Res. 2005;51:1–8. doi: 10.1016/j.neures.2004.09.007. [DOI] [PubMed] [Google Scholar]
  • 256.Nelson LE, Guo TZ, Lu J, Saper CB, Franks NP, Maze M. The sedative component of anesthesia is mediated by GABAA receptors in an endogenous sleep pathway. Nat. Neurosci. 2002;5:979–984. doi: 10.1038/nn913. [DOI] [PubMed] [Google Scholar]
  • 257.Nelson LE, Lu J, Guo TZ, Saper CB, Franks NP, Maze M. The α2-adrenoceptor agonist dexmedetomidine converges on an endogenous sleep-promoting pathway to exert its sedative effects. Anesthesiology. 2003;98:428– 436. doi: 10.1097/00000542-200302000-00024. [DOI] [PubMed] [Google Scholar]
  • 258.Nemoto T, Konno A, Chiba T. Synaptic contact of neuropeptide- and amine-containing axons on parasympathetic preganglionic neurons in the superior salivatory nucleus of the rat. Brain Res. 1995;685:33–45. doi: 10.1016/0006-8993(95)00409-j. [DOI] [PubMed] [Google Scholar]
  • 259.Nicholson JE, Severin CM. The superior and inferior salivatory nuclei in the rat. Neurosci. Lett. 1981;21:149–154. doi: 10.1016/0304-3940(81)90373-6. [DOI] [PubMed] [Google Scholar]
  • 260.Nicola SM, Taha SA, Kim SW, Fields HL. Nucleus accumbens dopamine release is necessary and sufficient to promote the behavioural response to reward-predictive cues. Neuroscience. 2005;135:1025–1033. doi: 10.1016/j.neuroscience.2005.06.088. [DOI] [PubMed] [Google Scholar]
  • 261.Nieuwenhuys R. Chemoarchitecture of the brain. Berlin: Springer- Verlag; 1985. [Google Scholar]
  • 262.Nishiike S, Takeda N, Kubo T, Nakamura S. Neurons in rostral ventrolateral medulla mediate vestibular inhibition of locus coeruleus in rats. Neuroscience. 1997;77:219–232. doi: 10.1016/s0306-4522(96)00436-8. [DOI] [PubMed] [Google Scholar]
  • 263.Nitz D, Siegel JM. GABA release in the locus coeruleus as a function of sleep/wake state. Neuroscience. 1997;78:795–801. doi: 10.1016/s0306-4522(96)00549-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Nosaka S, Yasunaga K, Tamai S. Vagal cardiac preganglionic neurons: distribution, cell types, and reflex discharges. Am. J. Physiol. 1982;243:R92–R98. doi: 10.1152/ajpregu.1982.243.1.R92. [DOI] [PubMed] [Google Scholar]
  • 265.Nygren L-G, Olson L. A new major projection from locus coeruleus: the main source of noradrenergic nerve terminals in the ventral and dorsal columns of the spinal cord. Brain Res. 1977;132:85–93. doi: 10.1016/0006-8993(77)90707-7. [DOI] [PubMed] [Google Scholar]
  • 266.Oades RD, Halliday GM. Ventral tegmental (A10) system: neurobiology 1 Anatomy and connectivity. Brain Res. 1987;434:117–165. doi: 10.1016/0165-0173(87)90011-7. [DOI] [PubMed] [Google Scholar]
  • 267.Olave MJ, Maxwell DJ. An investigation of neurones that possess the α2C-adrenergic receptor in the rat dorsal horn. Neuroscience. 2002;115:31–40. doi: 10.1016/s0306-4522(02)00407-4. [DOI] [PubMed] [Google Scholar]
  • 268.Olson L, Fuxe K. On the projections from the locus coeruleus noradrenaline neurons: the cerebellar innervation. Brain Res. 1971;28:165–171. doi: 10.1016/0006-8993(71)90533-6. [DOI] [PubMed] [Google Scholar]
  • 269.Ornstein K, Milon H, McRae-Degueurce A, Alvarez C, Berger B, Würzner HP. Biochemical and radioautographic evidence for dopaminergic afferents of the locus coeruleus originating in the ventral tegmental area. J. Neural Transm. 1987;70:183–191. doi: 10.1007/BF01253597. [DOI] [PubMed] [Google Scholar]
  • 270.Osaka T, Matsumura H. Noradrenergic inputs to sleep-related neurons in the preoptic area from the locus coeruleus and the ventrolateral medulla in the rat. Neurosci. Res. 1994;19:39–50. doi: 10.1016/0168-0102(94)90006-x. [DOI] [PubMed] [Google Scholar]
  • 271.Pan ZZ, Grudt TJ, Williams JT. Alpha 1-adrenoceptors in rat dorsal raphe neurons: regulation of two potassium conductances. J. Physiol. (Lond.) 1994;478:437–447. doi: 10.1113/jphysiol.1994.sp020263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Papay R, Gaivin R, Jha A, McCune DF, McGrath JC, Rodrigo MC, Simpson PC, Doze VA, Perez DM. Localization of the mouse α1A-adrenergic receptor (AR) in the brain: α1AAR is expressed in neurons, GABAergic interneurons, and NG2 oligodendrocyte progenitors. J. Comp. Neurol. 2006;497:209–222. doi: 10.1002/cne.20992. [DOI] [PubMed] [Google Scholar]
  • 273.Papay R, Gaivin R, McCune DF, Rorabaugh BR, Macklin WB, McGrath JC, Perez DM. Mouse α1B-adrenergic receptor is expressed in neurons and NG2 oligodendrocytes. J. Comp. Neurol. 2004;478:1–10. doi: 10.1002/cne.20215. [DOI] [PubMed] [Google Scholar]
  • 274.Parkis MA, Bayliss DA, Berger AJ. Actions of norepinephrine on rat hypoglossal motoneurons. J. Neuropsysiol. 1995;74:1911–1919. doi: 10.1152/jn.1995.74.5.1911. [DOI] [PubMed] [Google Scholar]
  • 275.Pascual J, del Arco C, Gonzalez AM, Pazos A. Quantitative light microscope autoradiographic localization of alpha 2-adrenoceptors in the human brain. Brain Res. 1992;585:116–127. doi: 10.1016/0006-8993(92)91196-l. [DOI] [PubMed] [Google Scholar]
  • 276.Pasquier DA, Kemper TL, Forbes WB, Morgane PJ. Dorsal raphe, substantia nigra and locus coeruleus: interconnections with each other and the neostriatum. Brain Res. Bull. 1977;2:323–339. doi: 10.1016/0361-9230(77)90066-1. [DOI] [PubMed] [Google Scholar]
  • 277.Pasquier DA, Reinoso-Suarez F. The topographic organization of hypothalamic and brain stem projections to the hippocampus. Brain Res. Bull. 1978;3:373–389. doi: 10.1016/0361-9230(78)90106-5. [DOI] [PubMed] [Google Scholar]
  • 278.Pepper CM, Henderson G. Opiates and opioid peptides hyperpolarize locus coeruleus neurons in vitro. Science. 1980;209:394–395. doi: 10.1126/science.7384811. [DOI] [PubMed] [Google Scholar]
  • 279.Peyron C, Luppi P-H, Rampon C, Jouvet M. Localization of the GABA-ergic neurons projecting to the rat dorsal raphe and locus coeruleus. Soc. Neurosci. Abs. 1995;21:373. [Google Scholar]
  • 280.Peyron C, Tighe DK, van den Pol AN, de Lecea L, Heller HC, Sutcliffe JG, Kilduff TS. Neurons containing hypocretin (orexin) project to multiple neuronal systems. J. Neurosci. 1998;18:9996–10015. doi: 10.1523/JNEUROSCI.18-23-09996.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 281.Pieribone VA, Nicholas AP, Dagerlind Å, Hökfelt T. Distribution of α1 adrenoceptors in rat brain revealed by in situ hybridization experiments utilizing subtype-specific probes. J. Neurosci. 1994;14:4252–4268. doi: 10.1523/JNEUROSCI.14-07-04252.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 282.Pillot C, Heron A, Cochins V, Tardivel-Lacombe J, Ligneau X, Schwartz J-C, Arrang J-M. A detailed mapping of the histamine H3 receptor and its gene transcripts in rat brain. Neuroscience. 2002;114:173–193. doi: 10.1016/s0306-4522(02)00135-5. [DOI] [PubMed] [Google Scholar]
  • 283.Proudfit HK, Clark FM. The projections of locus coeruleus neurons to the spinal cord. Prog. Brain Res. 1991;88:123–141. doi: 10.1016/s0079-6123(08)63803-0. [DOI] [PubMed] [Google Scholar]
  • 284.Pudovkina OL, Cremers TIFH, Westerink BHC. Regulation of the release of serotonin in the dorsal raphe nucleus by α1 and α2 adrenoceptors. Synapse. 2003;50:77–82. doi: 10.1002/syn.10245. [DOI] [PubMed] [Google Scholar]
  • 285.Pudovkina OL, Westerink BHC. Release of noradrenaline in the locus coeruleus. In: Ludwig M, editor. Dendritic Neurotransmitter Release. USA: Springer; 2005. pp. 145–154. [Google Scholar]
  • 286.Quintin L, Buda M, Hilaire G, Bardelay C, Ghignone M, Pujol JF. Catecholamine metabolism in the rat locus coeruleus as studied by in vivo differential pulse voltammetry. III. Evidence for the existence of an alpha 2-adrenergic tonic inhibition in behaving rats. Brain Res. 1986;375:235–245. doi: 10.1016/0006-8993(86)90743-2. [DOI] [PubMed] [Google Scholar]
  • 287.Rajkowski J, Kubiak P, Aston-Jones G. Locus coeruleus activity in monkey: phasic and tonic changes are associated with altered vigilance. Brain Res. Bull. 1994;35:607–616. doi: 10.1016/0361-9230(94)90175-9. [DOI] [PubMed] [Google Scholar]
  • 288.Rampon C, Peyron C, Gervasoni D, Pow DV, Luppi P-H, Fort P. Origins of the glycinergic inputs to the rat locus coeruleus and dorsal raphe nuclei: a study combining retrograde tracing with glycine immunohistochemistry. Eur. J. Neurosci. 1999;11:1058–1066. doi: 10.1046/j.1460-9568.1999.00511.x. [DOI] [PubMed] [Google Scholar]
  • 289.Rasmussen K, Aghajanian GK. Serotonin excitation of facial motoneurons: receptor subtype characterization. Synapse. 1990;5:324–332. doi: 10.1002/syn.890050409. [DOI] [PubMed] [Google Scholar]
  • 290.Rasmussen K, Jacobs BL. Single unit activity of locus coeruleus neurons in the freely moving cat. II. Conditioning and pharmacologic studies. Brain Res. 1986;23:335–344. doi: 10.1016/0006-8993(86)90371-9. [DOI] [PubMed] [Google Scholar]
  • 291.Rasmussen K, Morilak DA, Jacobs B. Single unit activity of locus coeruleus neurons in the freely moving cat. I. During naturalistic behaviors and in response to simple and complex stimuli. Brain Res. 1986;371:324–334. doi: 10.1016/0006-8993(86)90370-7. [DOI] [PubMed] [Google Scholar]
  • 292.Rathner JA, Owens NC, McAllen RM. Cold-activated raphe- spinal neurons in rats. J. Physiol. 2001;535:841–854. doi: 10.1111/j.1469-7793.2001.t01-1-00841.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 293.Redmond DE, Huang YH. New evidence for a locus coeruleus- norepinephrine connection with anxiety. Life Sci. 1979;25:2149–2162. doi: 10.1016/0024-3205(79)90087-0. [DOI] [PubMed] [Google Scholar]
  • 294.Redmond DE, Huang YH, Snyder DR, Maas JW. Behavioural effects of stimulation of the nucleus locus coeruleus in the stump-tailed monkey Macaca arctoides. Brain Res. 1976;116:502–510. doi: 10.1016/0006-8993(76)90498-4. [DOI] [PubMed] [Google Scholar]
  • 295.Reichlin S. In: Neuroendocrinology Williams Textbook of Endocrinology. Wilson JD, Foster DW, Kronenberg HM, Larsen PR, editors. Philadelphia: WB Saunders Company; 1998. pp. 165–248. [Google Scholar]
  • 296.Reyes BAS, Valentino RJ, Xu G, Van Bockstaele EJV. Hypothalamic projections to locus coeruleus neurons in rat brain. Eur. J. Neurosci. 2005;22:93–106. doi: 10.1111/j.1460-9568.2005.04197.x. [DOI] [PubMed] [Google Scholar]
  • 297.Rivot JP, Chiang CY, Besson JM. Increase of serotonin metabolism within the dorsal horn of the spinal cord during nucleus raphe magnus stimulation, as revealed by in vivo electrochemical detection. Brain Res. 1982;238:117–126. doi: 10.1016/0006-8993(82)90775-2. [DOI] [PubMed] [Google Scholar]
  • 298.Robbins TW. Cortical noradrenaline, attention and arousal. Psychol. Med. 1984;14:13–21. doi: 10.1017/s0033291700003032. [DOI] [PubMed] [Google Scholar]
  • 299.Robertson HA, Leslie RA. Noradrenergic alpha 2 binding sites in vagal dorsal motor nucleus and nucleus tractus solitarius: autoradiographic localization. Can. J. Physiol. Pharmacol. 1985;63:1190–1194. doi: 10.1139/y85-195. [DOI] [PubMed] [Google Scholar]
  • 300.Rosin DL, Talley EM, Lee A, Stornetta RL, Gaylinn BD, Guyenet PG, Lynch KR. Distribution of alpha 2C-adrenergic receptor-like immunoreactivity in the rat central nervous system. J. Comp. Neurol. 1996;372:135–165. doi: 10.1002/(SICI)1096-9861(19960812)372:1<135::AID-CNE9>3.0.CO;2-4. [DOI] [PubMed] [Google Scholar]
  • 301.Saigal RP, Karamanlidis AN, Voogd J, Michaloudi H, Mangana O. Cerebellar afferents from motor nuclei of cranial nerves, the nucleus of the solitary tract, and nuclei coeruleus and parabrachialis in sheep, demonstrated with retrograde transport of horseradish peroxidase. Brain Res. 1980;197:200–206. doi: 10.1016/0006-8993(80)90445-x. [DOI] [PubMed] [Google Scholar]
  • 302.Saint-Mleux B, Eggermann E, Bisetti A, Bayer L, Machard D, Jones BE, Mϋhlethaler M, Serafin M. Nicotinic enhancement of the noradrenergic inhibition of sleep-promoting neurons in the ventrolateral preoptic area. J. Neurosci. 2004;24:63–67. doi: 10.1523/JNEUROSCI.0232-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 303.Sakai K. Executive mechanisms of paradoxical sleep. Arch. Ital. Biol. 1988;126:239–257. [PubMed] [Google Scholar]
  • 304.Sakai K, Salvert D, Touret M, Jouvet M. Afferent connections of the nucleus raphe dorsalis in the cat as visualized by the horseradish peroxidase technique. Brain Res. 1977;137:11–35. doi: 10.1016/0006-8993(77)91010-1. [DOI] [PubMed] [Google Scholar]
  • 305.Samuels ER, Hou RH, Langley RW, Szabadi E, Bradshaw CM. Comparison of pramipexole and modafinil on arousal, autonomic, and endocrine functions in healthy volunteers. J. Psychopharmacol. 2006;20:756–770. doi: 10.1177/0269881106060770. [DOI] [PubMed] [Google Scholar]
  • 306.Samuels ER, Hou RH, Langley RW, Szabadi E, Bradshaw CM. Comparison of amisulpride and pramipexole on alertness, autonomic and endocrine functions in healthy volunteers. Psychopharmacology (Berl.) 2006;187:498–510. doi: 10.1007/s00213-006-0443-y. [DOI] [PubMed] [Google Scholar]
  • 307.Samuels ER, Hou RH, Langley RW, Szabadi E, Bradshaw CM. Modulation of the acoustic startle response by the level of arousal: comparison of clonidine and modafinil in healthy volunteers. Neuropsychopharmacology. 2007;32:2405–2421. doi: 10.1038/sj.npp.1301363. [DOI] [PubMed] [Google Scholar]
  • 308.Samuels ER, Hou RH, Langley RW, Szabadi E, Bradshaw CM. Comparison of pramipexole with and without domperidone co- administration on alertness, autonomic, and endocrine functions in healthy volunteers. Br. J. Clin. Pharmacol. 2007;64:591–602. doi: 10.1111/j.1365-2125.2007.02938.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 309.Sands SA, Morilak DA. Expression of α1D adrenergic receptor messenger RNA in oxytocin- and corticotropin-releasing hormone- synthesizing neurons in the rat paraventricular nucleus. Neuroscience. 1999;91:639– 649. doi: 10.1016/s0306-4522(98)00616-2. [DOI] [PubMed] [Google Scholar]
  • 310.Saper CB, Loewy AD, Swanson LW, Cowan WM. Direct hypothalamo-autonomic connections. Brain Res. 1976;117:305–312. doi: 10.1016/0006-8993(76)90738-1. [DOI] [PubMed] [Google Scholar]
  • 311.Sara SJ, Devauges V. Priming stimulation of locus coeruleus facilitates memory retrieval in the rat. Brain Res. 1988;438:299–303. doi: 10.1016/0006-8993(88)91351-0. [DOI] [PubMed] [Google Scholar]
  • 312.Sasa M, Munekiyo H, Ikeda S, Takaori S. Noradrenaline-mediated inhibition by locus coeruleus of spinal trigeminal neurons. Brain Res. 1974;80:443–460. doi: 10.1016/0006-8993(74)91029-4. [DOI] [PubMed] [Google Scholar]
  • 313.Sasa M, Ohno Y, Ito J, Kashii S, Utsumi S, Takaori S. Beta- receptor involvement in locus coeruleus-induced inhibition of spinal trigeminal nucleus neurons: microiontophoretic and HRP studies. Brain Res. 1986;377:337–343. doi: 10.1016/0006-8993(86)90877-2. [DOI] [PubMed] [Google Scholar]
  • 314.Sasa M, Takaori S. Influence of the locus coeruleus on transmission in the spinal trigeminal nucleus neurons. Brain Res. 1973;55:203–208. doi: 10.1016/0006-8993(73)90502-7. [DOI] [PubMed] [Google Scholar]
  • 315.Sasa M, Yoshimura N. Locus coeruleus noradrenergic neurons as a micturition center. Microsc. Res. Tech. 1994;29:226–230. doi: 10.1002/jemt.1070290308. [DOI] [PubMed] [Google Scholar]
  • 316.Satomi H, Yamamoto T, Ise H, Takahashi K. Identification of the inferior salivatory nucleus in the cat as studies by HRP bathings of the transected glossopharyngeal nerve root. Neurosci. Lett. 1979;11:259–263. doi: 10.1016/0304-3940(79)90004-1. [DOI] [PubMed] [Google Scholar]
  • 317.Sawchenko PE, Swanson LW. Immunohistochemical identification of neurons in the paraventricular nucleus of the hypothalamus that project to the medulla or to the spinal cord in the rat. J. Comp. Neurol. 1982;205:260–272. doi: 10.1002/cne.902050306. [DOI] [PubMed] [Google Scholar]
  • 318.Schambra UB, Mackensen GB, Stafford-Smith M, Haines DE, Schwinn DA. Neuron specific α-adrenergic receptor expression in human cerebellum: implications for emerging cerebellar roles in neurologic disease. Neuroscience. 2005;135:507–523. doi: 10.1016/j.neuroscience.2005.06.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Scheinin M, Lomasney JW, Hayden-Hixson DM, Schambra UB, Caron MG, Lefkowitz RJ, Fremeau RT Jr. Distribution of alpha 2-adrenergic receptor subtype gene expression in rat brain. Brain Res. Mol. Brain Res. 1994;21:133–149. doi: 10.1016/0169-328x(94)90386-7. [DOI] [PubMed] [Google Scholar]
  • 320.Schuerger RJ, Balaban CD. Immunohistochemical demonstration of regionally selective projections from locus coeruleus to the vestibular nuclei in rats. Exp. Brain Res. 1993;92:351–359. doi: 10.1007/BF00229022. [DOI] [PubMed] [Google Scholar]
  • 321.Séguéla P, Watkins KC, Geffard M, Descarries L. Noradrenaline axon terminals in adult rat neocortex: an immunocytochemical analysis in serial thin sections. Neuroscience. 1990;35:249–264. doi: 10.1016/0306-4522(90)90079-j. [DOI] [PubMed] [Google Scholar]
  • 322.Senba E, Tohyama M, Shiosaka H, Takagi M, Sakanaka T, Matuzaki Y, Takahashi N, Shimizu N. Experimental and morphological studies of the noradrenaline innervations in the nucleus tractus spinal nervi trigemini of the rat with special reference to their fine structures. Brain Res. 1981;206:39–50. doi: 10.1016/0006-8993(81)90099-8. [DOI] [PubMed] [Google Scholar]
  • 323.Shao YP, Sutin J. Noradrenergic facilitation of motor neurons: localization of adrenergic receptors in neurons and nonneuronal cells in the trigeminal motor nucleus. Exp. Neurol. 1991;114:216–227. doi: 10.1016/0014-4886(91)90038-e. [DOI] [PubMed] [Google Scholar]
  • 324.Sherin JE, Elmquist JK, Torrealba F, Saper CB. Innervation of histaminergic tuberomamillary neurons by GABAergic and galaninergic neurons in the ventrolateral preoptic nucleus of the rat. J. Neurosci. 1998;18:4705– 4721. doi: 10.1523/JNEUROSCI.18-12-04705.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 325.Shi TJ, Winzer-Serhan U, Leslie F, Hokfelt T. Distribution of alpha2-adrenoceptor mRNAs in the rat lumber spinal cord in normal and axotomized rats. Neuroreport. 1999;10:2835–2839. doi: 10.1097/00001756-199909090-00025. [DOI] [PubMed] [Google Scholar]
  • 326.Shih C-D, Chan SHH, Chan JYH. Participation of hypothalamic paraventricular nucleus in locus coeruleus-induced baroreflex suppression in rats. Am. J. Physiol. 1995;269:H46–H52. doi: 10.1152/ajpheart.1995.269.1.H46. [DOI] [PubMed] [Google Scholar]
  • 327.Siegel JM. The stuff dreams are made of: anatomical substrates of REM sleep. Nat. Neurosci. 2006;9:721–722. doi: 10.1038/nn0606-721. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 328.Sim LJ, Joseph SA. Efferent projections of the nucleus raphe magnus. Brain Res. Bull. 1992;28:679–682. doi: 10.1016/0361-9230(92)90246-t. [DOI] [PubMed] [Google Scholar]
  • 329.Sim LJ, Joseph SA. Dorsal raphe nucleus efferents: termination in peptidergic fields. Peptides. 1993;14:75–83. doi: 10.1016/0196-9781(93)90013-7. [DOI] [PubMed] [Google Scholar]
  • 330.Simon H, Le Moal M, Stinus L, Calas A. Anatomical relationships between the ventral tegmental mesencephalic tegmentum-a10 region and the locus coeruleus as demonstrated by anterograde and retrograde tracing techniques. J. Neural Transm. 1979;44:77–86. doi: 10.1007/BF01252703. [DOI] [PubMed] [Google Scholar]
  • 331.Simson PE. Blockade of alpha2-adrenergic receptors markedly potentiates glutamate-evoked activity of locus coeruleus neurons. Int. J. Neurosci. 2001;106:95–99. doi: 10.3109/00207450109149740. [DOI] [PubMed] [Google Scholar]
  • 332.Singewald N, Philippu A. Release of neurotransmitters in the locus coeruleus. Prog. Neurobiol. 1998;56:237–267. doi: 10.1016/s0301-0082(98)00039-2. [DOI] [PubMed] [Google Scholar]
  • 333.Singewald N, Sharp T. Neuroanatomical targets of anxiogenic drugs in the hindbrain as revealed by Fos immunochemistry. Neuroscience. 2000;98:759–770. doi: 10.1016/s0306-4522(00)00177-9. [DOI] [PubMed] [Google Scholar]
  • 334.Sluka KA, Westlund KN. Spinal projections of the locus coeruleus and the nucleus subcoeruleus in the Harlan and the Sasco Sprague-Dawley rat. Brain Res. 1992;579:67–73. doi: 10.1016/0006-8993(92)90742-r. [DOI] [PubMed] [Google Scholar]
  • 335.Smith A. Analgesia: Dial ‘P’ for pain. Nat. Rev. Neurosci. 2002;3:173. [Google Scholar]
  • 336.Smith MS, Schambra UB, Wilson KH, Page SO, Hulette C, Light AR, Schwinn DA. Alpha 2-adrenergic receptors in human spinal cord: specific localized expression of mRNA encoding alpha 2-adrenergic receptor subtypes at four distinct levels. Brain Res. Mol. Brain Res. 1995;34:109–117. doi: 10.1016/0169-328x(95)00148-l. [DOI] [PubMed] [Google Scholar]
  • 337.Smith MS, Schambra UB, Wilson KH, Page SO, Schwinn DA. α1-adrenergic receptors in human spinal cord: specific localized expression of mRNA encoding α1-adrenergic receptor subtypes at four distinct levels. Mol. Brain Res. 1999;63:254–261. doi: 10.1016/s0169-328x(98)00287-3. [DOI] [PubMed] [Google Scholar]
  • 338.Spencer SE, Sawyer WB, Wada H, Platt KB, Loewy AD. CNS projections to the pterygopalatine parasympathetic preganglionic neurons in the rat: a retrograde transneuronal viral cell body labeling study. Brain Res. 1990;534:149–169. doi: 10.1016/0006-8993(90)90125-u. [DOI] [PubMed] [Google Scholar]
  • 339.Steininger TL, Gong H, McGinty D, Szymusiak R. Subregional organization of preoptic area/anterior hypothalamic projections to arousal- related monoaminergic cell groups. J. Comp. Neurol. 2001;429:638–653. [PubMed] [Google Scholar]
  • 340.Sterpenich V, D’Argembeau A, Desseilles M, Balteau E, Albouy G, Vandewalle G, Degueldre C, Luxen A, Collette F, Maquet P. The locus coeruleus is involved in the successful retrieval of emotional memories in humans. J. Neurosci. 2006;26:7416–7423. doi: 10.1523/JNEUROSCI.1001-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 341.Stock G, Rupprecht U, Stumpf H, Schlor KH. Cardiovascular changes during arousal elicited by stimulation of amygdala, hypothalamus and locus coeruleus. J. Auton. Nerv. Syst. 1981;3:503–510. doi: 10.1016/0165-1838(81)90083-7. [DOI] [PubMed] [Google Scholar]
  • 342.Stone EA, Quartermain D, Lin Y, Lehmann ML. Central α1- adrenergic system in behavioural activity and depression. Biochem. Pharmacol. 2006;73:1063–1075. doi: 10.1016/j.bcp.2006.10.001. [DOI] [PubMed] [Google Scholar]
  • 343.Sullivan RM, Wilson DA, Lemon C, Gerhardt GA. Bilateral 6- OHDA lesions of the locus coeruleus impair associative olfactory learning in newborn rats. Brain Res. 1994;643:306–309. doi: 10.1016/0006-8993(94)90038-8. [DOI] [PubMed] [Google Scholar]
  • 344.Sun M-K. Central neural organization and control of sympathetic nervous system in mammals. Prog. Neurobiol. 1995;47:157–233. doi: 10.1016/0301-0082(95)00026-8. [DOI] [PubMed] [Google Scholar]
  • 345.Sun M-K, Guyenet PG. GABA-mediated baroreceptor inhibition of reticulospinal neurons. Am. J. Physiol. 1985;249:R672–680. doi: 10.1152/ajpregu.1985.249.6.R672. [DOI] [PubMed] [Google Scholar]
  • 346.Sun M-K, Guyenet PG. Hypothalamic glutamatergic input to medullary sympathoexcitatory neurons in rats. Am. J. Physiol. 1986;251:R798–R810. doi: 10.1152/ajpregu.1986.251.4.R798. [DOI] [PubMed] [Google Scholar]
  • 347.Sutcliffe JG, de Lecea L. The hypocretins: setting the arousal threshold. Nat. Rev. Neurosci. 2002;3:339–349. doi: 10.1038/nrn808. [DOI] [PubMed] [Google Scholar]
  • 348.Suzuki M, Hurd YL, Sokoloff P, Schwartz J-C, Sedvall G. D3 dopamine receptor mRNA is widely expressed in the human brain. Brain Res. 1998;779:58–74. doi: 10.1016/s0006-8993(97)01078-0. [DOI] [PubMed] [Google Scholar]
  • 349.Swanson LW. The projections of the ventral tegmental area and adjacent regions: a combined fluorescent retrograde tracer and immunofluorescence study in the rat. Brain Res. Bull. 1982;9:321–353. doi: 10.1016/0361-9230(82)90145-9. [DOI] [PubMed] [Google Scholar]
  • 350.Swanson LW, Sawchenko PE. Paraventricular nucleus: a site for the integration of neuroendocrine and autonomic mechanisms. Neuroendocrinology. 1980;31:410–417. doi: 10.1159/000123111. [DOI] [PubMed] [Google Scholar]
  • 351.Swanson LW, Sawchenko PE. Hypothalamic integration: organization of the paraventricular and supraoptic nuclei. Ann. Rev. Neurosci. 1983;6:269–324. doi: 10.1146/annurev.ne.06.030183.001413. [DOI] [PubMed] [Google Scholar]
  • 352.Szabadi E, Bradshaw CM. Autonomic pharmacology of α2- adrenoceptors. J. Psychopharmacol. 1996;10(suppl 3):6–18. [Google Scholar]
  • 353.Szabadi E, Tavernor S. Hypo- and hypersalivation induced by psychoactive drugs. Incidence, mechanisms and therapeutic implications. CNS Drugs. 1999;11:449–466. [Google Scholar]
  • 354.Szymusiak R, Alam N, Steininger TL, McGinty D. Sleep-waking discharge patterns of ventrolateral preoptic/anterior hypothalamic neurons in rats. Brain Res. 1998;803:178–188. doi: 10.1016/s0006-8993(98)00631-3. [DOI] [PubMed] [Google Scholar]
  • 355.Takakura ACT, Moreira TS, De Luca LA Jr, Renzi A, Menani JV. Central α2 adrenergic receptors and cholinergic-induced salivation in rats. Brain Res. Bull. 2003;59:383–386. doi: 10.1016/s0361-9230(02)00929-2. [DOI] [PubMed] [Google Scholar]
  • 356.Talley EM, Rosin DL, Lee A, Guyenet PG, Lynch KR. Distribution of alpha 2A-adrenergic receptor-like immunoreactivity in the rat central nervous system. J. Comp. Neurol. 1996;372:111–134. doi: 10.1002/(SICI)1096-9861(19960812)372:1<111::AID-CNE8>3.0.CO;2-6. [DOI] [PubMed] [Google Scholar]
  • 357.Tanaka M, Yoshida M, Emoto H, Ishii H. Noradrenaline systems in the hypothalamus, amygdala and locus coeruleus are involved in the provocation of anxiety: basic studies. Eur. J. Pharmacol. 2000;405:397–406. doi: 10.1016/s0014-2999(00)00569-0. [DOI] [PubMed] [Google Scholar]
  • 358.Tavares A, Handy DE, Bogdanova NN, Rosene DL, Gavras H. Localization of α2A- and α2B-adrenergic receptor subtypes in brain. Hypertension. 1996;27:449–455. doi: 10.1161/01.hyp.27.3.449. [DOI] [PubMed] [Google Scholar]
  • 359.Ter Horst GJ, Toes GJ, Van Willigen JD. Locus coeruleus projections to the dorsal motor vagus nucleus in the rat. Neuroscience. 1991;45:153–160. doi: 10.1016/0306-4522(91)90111-z. [DOI] [PubMed] [Google Scholar]
  • 360.Thomas DN, Nutt D, Holman RB. Regionally specific changes in extracellular noradrenaline following chronic idazoxan as revealed by in vivo microdialysis. Eur. J. Pharmacol. 1994;261:53–57. doi: 10.1016/0014-2999(94)90299-2. [DOI] [PubMed] [Google Scholar]
  • 361.Thompson AM. Pontine sources of norepinephrine in the cat cochlear nucleus. J. Comp. Neurol. 2003;457:374–383. doi: 10.1002/cne.10540. [DOI] [PubMed] [Google Scholar]
  • 362.Thorpy M. Current concepts in the etiology, diagnosis and treatment of narcolepsy. Sleep Med. 2001;2:5–17. doi: 10.1016/s1389-9457(00)00081-2. [DOI] [PubMed] [Google Scholar]
  • 363.Thrivikraman KV, Plotsky PM, Gann DS. Alterations of locus coeruleus noradrenergic activity in relation to pituitary secretion after hemorrhage in cats. Neurosci. Lett. 1993;161:85–88. doi: 10.1016/0304-3940(93)90146-c. [DOI] [PubMed] [Google Scholar]
  • 364.Tohyama M, Maeda T, Shimizu N. Detailed noradrenaline pathways of locus coeruleus neuron to the cerebral cortex with use of 6-hydroxydopa. Brain Res. 1974;79:139–144. doi: 10.1016/0006-8993(74)90573-3. [DOI] [PubMed] [Google Scholar]
  • 365.Tóth IE, Boldogkői Z, Medveczky I, Palkovits M. Lacrimal preganglionic neurons form a subdivision of the superior salivatory nucleus of rat: transneuronal labelling by pseudorabies virus. J. Auton. Nerv. Syst. 1999;77:45–54. doi: 10.1016/s0165-1838(99)00032-6. [DOI] [PubMed] [Google Scholar]
  • 366.Trulson ME, Jacobs BL. Raphe unit activity in freely moving cats: correlation with level of behavioural arousal. Brain Res. 1979;163:135–150. doi: 10.1016/0006-8993(79)90157-4. [DOI] [PubMed] [Google Scholar]
  • 367.Tsuruoka M, Matsutani K, Maeda M, Inoue T. Coeruleotrigeminal inhibition of nociceptive processing in the rat trigeminal subnucleus caudalis. Brain Res. 2003;993:146–153. doi: 10.1016/j.brainres.2003.09.023. [DOI] [PubMed] [Google Scholar]
  • 368.Uhde TW, Boulenger JP, Post RM, Siever LJ, Vittone BJ, Jimerson DC, Roy-Byrne PP. Fear and anxiety: relationship to noradrenergic function. Psychopathology. 1984;17(Suppl 3):8–23. doi: 10.1159/000284127. [DOI] [PubMed] [Google Scholar]
  • 369.Ungerstedt U. Stereotaxic mapping of the monoamine pathways in the rat brain. Acta Physiol. Scand. Suppl. 1971;367:1–49. doi: 10.1111/j.1365-201x.1971.tb10998.x. [DOI] [PubMed] [Google Scholar]
  • 370.Unnerstall JR, Kopajtic TA, Kuhar MJ. Distribution of alpha 2 agonist binding sites in the rat and human central nervous system: analysis of some functional, anatomic correlates of the pharmacologic effects of clonidine and related adrenergic agents. Brain Res. 1984;319:69–101. doi: 10.1016/0165-0173(84)90030-4. [DOI] [PubMed] [Google Scholar]
  • 371.Valentino RJ, Wehby RG. Morphine effects on locus coeruleus neurons are dependent on the state of arousal and availability of external stimuli: studies in anesthetized and unanesthetized rats. J. Pharmacol. Exp. Ther. 1988;244:1178–1186. [PubMed] [Google Scholar]
  • 372.Van Bockstaele EJ. Morphological substrates underlying opioid, epinephrine and γ-aminobutyric acid inhibitory actions in the rat locus coeruleus. Brain Res. Bull. 1998;47:1–15. doi: 10.1016/s0361-9230(98)00062-8. [DOI] [PubMed] [Google Scholar]
  • 373.Van Bockstaele EJ, Aston-Jones G. Collateralized projections from neurons in the rostral medulla to the nucleus locus coeruleus, the nucleus of the solitary tract and the periaqueductal gray. Neuroscience. 1992;49:653–668. doi: 10.1016/0306-4522(92)90234-s. [DOI] [PubMed] [Google Scholar]
  • 374.Van Bockstaele EJ, Colago EE, Moriwaki A, Uhl GR. Mu- opioid receptor is located on the plasma membrane of dendrites that receive asymmetric synapses from axon terminals containing leucine-enkephalin in the rat nucleus locus coeruleus. J. Comp. Neurol. 1996;376:65–74. doi: 10.1002/(SICI)1096-9861(19961202)376:1<65::AID-CNE4>3.0.CO;2-M. [DOI] [PubMed] [Google Scholar]
  • 375.Van Bockstaele EJ, Colago EEO, Valentino RJ. Amygdaloid corticotrophin-releasing factor targets locus coeruleus dendrites: substrate for the co-ordination of emotional and cognitive limbs of the stress response. J. Neurosci. 1998;10:743–757. doi: 10.1046/j.1365-2826.1998.00254.x. [DOI] [PubMed] [Google Scholar]
  • 376.Van Bockstaele EJ, Pieribone VA, Aston-Jones G. Diverse afferents converge on the nucleus paragigantocellularis in the rat ventrolateral medulla: retrograde and anterograde tracing studies. J. Comp. Neurol. 1989;290:561–584. doi: 10.1002/cne.902900410. [DOI] [PubMed] [Google Scholar]
  • 377.Van den Pol AN. The magnocellular and parvocellular paraventricular nucleus of rat: intrinsic organization. J. Comp. Neurol. 1982;206:317–345. doi: 10.1002/cne.902060402. [DOI] [PubMed] [Google Scholar]
  • 378.VanderMaelen CP, Aghajanian GK. Intracellular studies showing modulation of facial motoneurone excitability by serotonin. Nature. 1980;287:346–347. doi: 10.1038/287346a0. [DOI] [PubMed] [Google Scholar]
  • 379.Van Dongen PA. The central noradrenergic transmission and the locus coeruleus: a review of the data, and their implications for neurotransmission and neuromodulation. Prog. Neurobiol. 1981;16:117–143. doi: 10.1016/0301-0082(81)90009-5. [DOI] [PubMed] [Google Scholar]
  • 380.Van Gaalen M, Kawahara H, Kawahara Y, Westerink BHC. The locus coeruleus noradrenergic system in the rat brain studied by dual-probe microdialysis. Brain Res. 1997;763:56–62. doi: 10.1016/s0006-8993(97)00416-2. [DOI] [PubMed] [Google Scholar]
  • 381.Verret L, Fort P, Gervasoni D, Léger L, Luppi P-H. Localization of the neurons active during paradoxical (REM) sleep and projecting to the locus coeruleus noradrenergic neurons in the rat. J. Comp. Neurol. 2006;495:573–586. doi: 10.1002/cne.20891. [DOI] [PubMed] [Google Scholar]
  • 382.Vertes RP, Kocsis B. Projections of the dorsal raphe nucleus to the brainstem: PHA-L analysis in the rat. J. Comp. Neurol. 1994;340:11–26. doi: 10.1002/cne.903400103. [DOI] [PubMed] [Google Scholar]
  • 383.Volgin DV, Mackiewicz M, Kubin L. α1B receptors are the main postsynaptic mediators of adrenergic excitation in brainstem motoneurons, a single-cell RT-PCR study. J. Chem. Neuroanat. 2001;22:157–166. doi: 10.1016/s0891-0618(01)00124-7. [DOI] [PubMed] [Google Scholar]
  • 384.Wallace DM, Magnuson DJ, Gray TS. The amygdalo-brainstem pathway: selective innervation of dopaminergic, noradrenergic and adrenergic cells in the rat. Neurosci. Lett. 1989;97:252–258. doi: 10.1016/0304-3940(89)90606-x. [DOI] [PubMed] [Google Scholar]
  • 385.Ward DG, Gunn CG. Locus coeruleus complex: elicitation of a pressor response and a brain stem region necessary for its occurrence. Brain Res. 1976;107:401–406. doi: 10.1016/0006-8993(76)90236-5. [DOI] [PubMed] [Google Scholar]
  • 386.Watson M, McElligott JG. Cerebellar noradrenaline depletion and impaired acquisition of specific locomotor tasks in rats. Brain Res. 1984;296:129–138. doi: 10.1016/0006-8993(84)90518-3. [DOI] [PubMed] [Google Scholar]
  • 387.Westlund KN, Bowker RM, Ziegler MG, Coulter JD. Noradrenergic projections to the spinal cord of the rat. Brain Res. 1983;263:15–31. doi: 10.1016/0006-8993(83)91196-4. [DOI] [PubMed] [Google Scholar]
  • 388.Westlund KN, Coulter JD. Descending projections of the locus coeruleus and subcoeruleus/medial parabrachial nuclei in monkey: axonal transport studies and dopamine-beta-hydroxylase immunocytochemistry. Brain Res. 1980;2:235–264. doi: 10.1016/0165-0173(80)90009-0. [DOI] [PubMed] [Google Scholar]
  • 389.Westlund KN, Zhang D, Carlton SM, Sorkin LS, Willis WD. Noradrenergic innervation of somatosensory thalamus and spinal cord. Prog. Brain Res. 1991;88:77–88. doi: 10.1016/s0079-6123(08)63800-5. [DOI] [PubMed] [Google Scholar]
  • 390.White SR, Fung SJ, Barnes CD. Norepinephrine effects on spinal motoneurons. Prog. Brain Res. 1991;88:343–350. doi: 10.1016/s0079-6123(08)63821-2. [DOI] [PubMed] [Google Scholar]
  • 391.Willett CJ, Rutherford JG, Gwyn DJ, Leslie RA. Projections between the hypothalamus and the dorsal vagal complex in the cat: an HRP and autoradiographic study. Brain Res. Bull. 1987;18:63–71. doi: 10.1016/0361-9230(87)90034-7. [DOI] [PubMed] [Google Scholar]
  • 392.Willette RN, Punnen-Grandy S, Krieger AJ, Sapru HN. Differential regulation of regional vascular resistance by the rostral and caudal ventrolateral medulla in the rat. J. Auton. Nerv. Syst. 1987;18:143–151. doi: 10.1016/0165-1838(87)90101-9. [DOI] [PubMed] [Google Scholar]
  • 393.Williams JT, Henderson G, North RA. Characterization of alpha 2- adrenoceptors which increase potassium conductance in rat locus coeruleus neurones. Neuroscience. 1985;14:95–101. doi: 10.1016/0306-4522(85)90166-6. [DOI] [PubMed] [Google Scholar]
  • 394.Williams JT, North RA. catecholamine inhibition of calcium action potentials in rat locus coeruleus neurones. Neuroscience. 1985;14:103–109. doi: 10.1016/0306-4522(85)90167-8. [DOI] [PubMed] [Google Scholar]
  • 395.Wise RA. Dopamine, learning and motivation. Nat. Rev. Neurosci. 2004;5:483–494. doi: 10.1038/nrn1406. [DOI] [PubMed] [Google Scholar]
  • 396.Wu M-F, Gulyani SA, Yau E, Mignot E, Phan B, Siegel JM. Locus coeruleus neurons: cessation of activity during cataplexy. Neuroscience. 1999;91:1389–1399. doi: 10.1016/s0306-4522(98)00600-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 397.Yaïci E-D, Rampin O, Calas A, Jestin A, McKenna KE, Leclerc P, Benoit G, Giuliano F. α2a and α2c adrenoceptors on spinal neurons controlling penile erection. Neuroscience. 2002;114:945–960. doi: 10.1016/s0306-4522(02)00367-6. [DOI] [PubMed] [Google Scholar]
  • 398.Yamanaka A, Muraki Y, Tsujino N, Goto K, Sakurai T. Regulation of orexin neurons by the monoaminergic and cholinergic systems. Biochem. Biophys. Res. Commun. 2003;303:120–129. doi: 10.1016/s0006-291x(03)00299-7. [DOI] [PubMed] [Google Scholar]
  • 399.Yamanaka A, Tsujino N, Funahashi H, Honda K, Guan J-L, Wang Q- P, Tominaga M, Goto K, Shioda S, Sakurai T. Orexins activate histaminergic neurons via the orexin 2 receptor. Biochem. Biophys. Res. Commun. 2002;290:1237–1245. doi: 10.1006/bbrc.2001.6318. [DOI] [PubMed] [Google Scholar]
  • 400.Yamazato M, Sakima A, Nakazato J, Sesoko S, Muratani H, Fukiyama K. Hypotensive and sedative effects of clonidine injected into the rostral ventrolateral medulla of conscious rats. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2001;281:1868–1876. doi: 10.1152/ajpregu.2001.281.6.R1868. [DOI] [PubMed] [Google Scholar]
  • 401.Yokoyama C, Okamura H, Nakajima T, Taguchi J, Ibata Y. Autoradiographic distribution of [3H]YM-09151-2, a high-affinity and selective antagonist ligand for the dopamine D2 receptor group, in the rat brain and spinal cord. J. Comp. Neurol. 1994;344:121–136. doi: 10.1002/cne.903440109. [DOI] [PubMed] [Google Scholar]
  • 402.Yoshimura M, Higashi H, Nishi S. Noradrenaline mediates slow excitatory synaptic potentials in rat dorsal raphe neurons in vitro. Neurosci. Lett. 1985;61:305–310. doi: 10.1016/0304-3940(85)90481-1. [DOI] [PubMed] [Google Scholar]
  • 403.Yoshimura N, Sasa M, Yoshida O, Takaori S. Alpha 1-adrenergic receptor-mediated excitation from the locus coeruleus of the sacral parasympathetic preganglionic neuron. Life Sci. 1990;47:789–797. doi: 10.1016/0024-3205(90)90551-2. [DOI] [PubMed] [Google Scholar]
  • 404.Young WS, Kuhar MJ. Noradrenergic α1 and α2 receptors: light microscopic autoradiographic localization. Proc. Natl. Acad. Sci. USA. 1980;77:1696–1700. doi: 10.1073/pnas.77.3.1696. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 405.Yu WH, Srinivasan R. origin of the preganglionic visceral efferent fibers to the glands in the rat tongue as demonstrated by the horseradish peroxidase method. Neurosci. Lett. 1980;19:143–148. doi: 10.1016/0304-3940(80)90185-8. [DOI] [PubMed] [Google Scholar]
  • 406.Zaborszky L, Cullinan WE, Luine VN. Catecholamine-cholinergic interaction in the basal forebrain. Prog. Brain Res. 1993;98:31–49. doi: 10.1016/s0079-6123(08)62379-1. [DOI] [PubMed] [Google Scholar]
  • 407.Zagon A, Smith AD. Monosynaptic projections to the rostral ventrolateral medulla oblongata to identified sympathetic preganglionic neurons. Neuroscience. 1993;54:729–743. doi: 10.1016/0306-4522(93)90243-9. [DOI] [PubMed] [Google Scholar]
  • 408.Zaretsky DV, Zaretskaia MV, DiMicco JA. Stimulation and blockade of GABAA receptors in the raphe pallidus: effects on body temperature, heart rate, and blood pressure in conscious rats. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2003;285:R110–R116. doi: 10.1152/ajpregu.00016.2003. [DOI] [PubMed] [Google Scholar]
  • 409.Zhang C, Guo Y-Q, Qiao J-T, Dafny N. Locus coeruleus modulates thalamic nociceptive responses via adrenoceptors. Brain Res. 1998;784:116–122. doi: 10.1016/s0006-8993(97)01197-9. [DOI] [PubMed] [Google Scholar]
  • 410.Zhang K-M, Wang X-M, Peterson AM, Chen W-Y, Mokha SS. α2-adrenoceptors modulate NMDA-evoked responses of neurons in superficial and deeper dorsal horn of the medulla. J. Neurophysiol. 1998;80:2210– 2214. doi: 10.1152/jn.1998.80.4.2210. [DOI] [PubMed] [Google Scholar]
  • 411.Zheng JQ, Seki M, Hayakawa T, Ito H, Zyo K. Descending projections from the paraventricular hypothalamic nucleus to the spinal cord: anterograde tracing study in the rat. Okajimas Folia Anat. Jpn. 1995;72:119–135. doi: 10.2535/ofaj1936.72.2-3_119. [DOI] [PubMed] [Google Scholar]

Articles from Current Neuropharmacology are provided here courtesy of Bentham Science Publishers

RESOURCES