untitled

37
http://jop.sagepub.com/ Journal of Psychopharmacology http://jop.sagepub.com/content/27/8/659 The online version of this article can be found at: DOI: 10.1177/0269881113490326 2013 27: 659 originally published online 12 June 2013 J Psychopharmacol Elemer Szabadi Functional neuroanatomy of the central noradrenergic system Published by: http://www.sagepublications.com On behalf of: British Association for Psychopharmacology can be found at: Journal of Psychopharmacology Additional services and information for http://jop.sagepub.com/cgi/alerts Email Alerts: http://jop.sagepub.com/subscriptions Subscriptions: http://www.sagepub.com/journalsReprints.nav Reprints: http://www.sagepub.com/journalsPermissions.nav Permissions: What is This? - Jun 12, 2013 OnlineFirst Version of Record - Jul 16, 2013 Version of Record >> at Queens University on February 12, 2014 jop.sagepub.com Downloaded from at Queens University on February 12, 2014 jop.sagepub.com Downloaded from at Queens University on February 12, 2014 jop.sagepub.com Downloaded from

Upload: api-247539984

Post on 21-Oct-2015

6 views

Category:

Documents


2 download

TRANSCRIPT

http://jop.sagepub.com/Journal of Psychopharmacology

http://jop.sagepub.com/content/27/8/659The online version of this article can be found at:

 DOI: 10.1177/0269881113490326 2013 27: 659 originally published online 12 June 2013J Psychopharmacol

Elemer SzabadiFunctional neuroanatomy of the central noradrenergic system

  

Published by:

http://www.sagepublications.com

On behalf of: 

  British Association for Psychopharmacology

can be found at:Journal of PsychopharmacologyAdditional services and information for    

  http://jop.sagepub.com/cgi/alertsEmail Alerts:

 

http://jop.sagepub.com/subscriptionsSubscriptions:  

http://www.sagepub.com/journalsReprints.navReprints:  

http://www.sagepub.com/journalsPermissions.navPermissions:  

What is This? 

- Jun 12, 2013OnlineFirst Version of Record  

- Jul 16, 2013Version of Record >>

at Queens University on February 12, 2014jop.sagepub.comDownloaded from at Queens University on February 12, 2014jop.sagepub.comDownloaded from at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Journal of Psychopharmacology27(8) 659 –693

© The Author(s) 2013Reprints and permissions: sagepub.co.uk/journalsPermissions.navDOI: 10.1177/0269881113490326jop.sagepub.com

Contents1. Introduction2. Noradrenergic nuclei3. The noradrenergic neurone4. Efferents of the locus coeruleus4.1 Telencephalon4.1.1 Neocortex4.1.2 Basal forebrain4.2 Limbic system4.2.1 Amygdala4.2.2 Hippocampus4.3 Diencephalon4.3.1 Thalamus4.3.2 Hypothalamus4.3.2.1 Ventrolateral preoptic area4.3.2.2 Lateral hypothalamic/perifornical area

4.3.2.3 Paraventricular nucleus4.3.2.4 Arcuate nucleus/tuberinfundibular area4.4 Brainstem4.4.1 Sympathetic premotor nuclei4.4.1.1 Rostroventrolateral medulla4.4.1.2 Caudal raphe nuclei

Functional neuroanatomy of the central noradrenergic system

Elemer Szabadi

AbstractThe central noradrenergic neurone, like the peripheral sympathetic neurone, is characterized by a diffusely arborizing terminal axonal network. The central neurones aggregate in distinct brainstem nuclei, of which the locus coeruleus (LC) is the most prominent. LC neurones project widely to most areas of the neuraxis, where they mediate dual effects: neuronal excitation by α1-adrenoceptors and inhibition by α2-adrenoceptors. The LC plays an important role in physiological regulatory networks. In the sleep/arousal network the LC promotes wakefulness, via excitatory projections to the cerebral cortex and other wakefulness-promoting nuclei, and inhibitory projections to sleep-promoting nuclei. The LC, together with other pontine noradrenergic nuclei, modulates autonomic functions by excitatory projections to preganglionic sympathetic, and inhibitory projections to preganglionic parasympathetic neurones. The LC also modulates the acute effects of light on physiological functions (‘photomodulation‘): stimulation of arousal and sympathetic activity by light via the LC opposes the inhibitory effects of light mediated by the ventrolateral preoptic nucleus on arousal and by the paraventricular nucleus on sympathetic activity. Photostimulation of arousal by light via the LC may enable diurnal animals to function during daytime. LC neurones degenerate early and progressively in Parkinson‘s disease and Alzheimer‘s disease, leading to cognitive impairment, depression and sleep disturbance.

KeywordsLocus coeruleus, noradrenaline, neuroanatomy, forebrain, telencephalon, diencephalon, brainstem, spinal cord, arousal, autonomic functions, photomodulation

AbbreviationsAD: Alzheimer‘s disease; BF: basal forebrain, CNS: central nervous system; CR: caudal raphe; CRF: corticotrophin releasing factor; DMH: dorsomedial hypothalamus; DMV: dorsal motor nucleus of the vagus; DR: dorsal raphe nucleus; EEG: electroencephalogram; EMG: electromyogram; EWN: Edinger-Westphal nucleus; GABA: γ-aminobutyric acid; GH: growth hormone; GHRH: growth hormone releasing hormone; 5HT: 5hydroxytryptamine; IML: intermediolateral cell column (of the spinal cord); ipRGC: intrinsically photosensitive retinal ganglion cell; LC: locus coeruleus; LDT: laterodorsal tegmental nucleus; LH/PF: lateral hypothalamic/perifornical area; NIF: non-image forming; NMDA: N-methyl D-aspartate; OPN: olivary pretectal nucleus; PAG: periaqueductal grey matter; PD: Parkinson‘s disease; PGi: nucleus paragigantocellularis lateralis; PPT: pedunculopontine tegmental nucleus; PrH: nucleus prepositus hypoglossi; PVN: paraventricular nucleus; REM: rapid eye movement; RVLM: rostroventrolateral medulla; SCN: suprahiasmatic nucleus; SWS: slow wave sleep; TMN: tuberomamillary nucleus; VLPO: ventrolateral preoptic area; VTA: ventral tegmental area.

Division of Psychiatry, University of Nottingham, Nottingham, UK

Corresponding author:Elemer Szabadi, Division of Psychiatry, School of Community Health Sciences, University of Nottingham, Room B109, Medical School, Queen‘s Medical Centre, Nottingham NG7 2UH, UK. Email: [email protected]

490326 JOP27810.1177/0269881113490326Journal of Psychopharmacology Szabadi2013

Review

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

660 Journal of Psychopharmacology 27(8)

4.4.2 Parasympathetic preganglionic nuclei4.4.2.1 Edinger-Westphal nucleus4.4.2.2 Salivatory nuclei4.4.2.3 Vagal nuclei4.4.3 Motor nuclei4.4.3.1 Oculomotor nucleus4.4.3.2 Trigeminal motor nucleus4.4.3.3 Facial nucleus4.4.3.4 Hypoglossal nucleus4.4.4 Sensory nuclei4.4.4.1 Sensory trigeminal nuclei4.4.4.2 Cochlear nuclei4.4.5 Sleep/arousal regulating nuclei4.4.5.1 Dorsal raphe nucleus4.4.5.2 Pedunculopontine tegmental and laterodorsal tegmental

nuclei4.5 Cerebellum4.6 Spinal cord4.6.1 Dorsal horn4.6.2 Ventral horn4.6.3. Intermediolateral cell column5. Afferents of the locus coeruleus5.1 Neocortex5.2 Amygdala5.3 Hypothalamus5.3.1 Ventrolateral preoptic area5.3.2 Lateral hypothalamic/perifornical area5.3.3 Tuberomammillary nucleus5.3.4 Paraventricular nucleus5.4 Brainstem5.4.1 Midbrain and pons5.4.1.1 Ventral tegmental area5.4.1.2 Periaqueductal grey matter5.4.1.3 Dorsal raphe nucleus5.4.1.4 Pedunculopontine and laterodorsal tegmental nuclei5.4.2 Medulla oblongata5.4.2.1 Rostral ventrolateral medulla5.4.2.2 Dorsomedial rostral medulla5.4.3 Sensory trigeminal nuclei5.5 Spinal cord6 The locus coeruleus in regulatory networks6.1 Sleep/arousal network6.2 Autonomic network6.3 Photomodulation network7 Neuropathology of the locus coeruleus8 Pathophysiology of the locus coeruleus9 Conclusions

1. IntroductionInterest in the noradrenergic system goes back to the discovery by Otto Loewi (1921) that a chemical substance released by sympa-thetic nerve endings transmits the effect of sympathetic stimula-tion to the heart. Although it was established relatively early that this substance was a catecholamine, probably adrenaline, it took another 30 years to accumulate a convincing body of evidence to show that it was noradrenaline (Von Euler, 1951). Soon after this, noradrenaline was found in the brain (Vogt, 1954), and later local-ized in distinct sets of neurones (Dahsltröm and Fuxe, 1964). For

a historical review of noradrenergic neurotransmission, the reader is referred to Iversen (1967) and Livett (1973).

The topic of this review is the functional neuroanatomy of the central noradrenergic system. After a brief description of the noradrenergic neurone, the efferent and afferent connections of the locus coeruleus (LC), the major noradrenergic nucleus, will be discussed. In this section, the Jacksonian concept of the hier-archical organization of the brain (Greenblatt, 1999) will be fol-lowed, starting at the cerebral cortex and finishing with the spinal cord (for examples, see Szabadi (1993)). Finally, the func-tional importance of these connections will be illustrated by con-sidering the role of the LC in a number of regulatory networks of the brain.

Due to limitations of space, this review is not aimed at com-prehensive coverage, but rather attempts to illustrate the func-tional importance of the anatomical hardware. A number of excellent reviews of the anatomy of the central noradrenergic system have been published over the years, covering different aspects of the topic at variable degrees of details (e.g. Amaral and Sinnamon, 1977; Aston-Jones et al., 1991b; Byrum and Guyenet, 1987; Cederbaum and Aghajnaian, 1978; Foote et al., 1983; Guyenet, 1991; Loughlin et al., 1986a, 1986b; Luppi et al., 1995; Moore and Bloom, 1979; Nieuwenhuys, 1985; Van Dongen, 1981). The present paper relies, in parts, on a recent review with a focus on the role of the LC in the regulation of arousal and autonomic functions (Samuels and Szabadi, 2008a, 2008b).

2. The noradrenergic neuroneThe nerve terminals of the peripheral noradrenergic sympathetic neurone arborize profusely to create a dense meshwork (‘sympa-thetic ground plexus‘). The terminal fibers contain enlargements (‘varicosities‘) which give a string-of-beads-like appearance to the fibers. The varicosities have the structure and function of pre-synaptic terminals or ‘boutons‘, and have been identified as sites of release (Geffen and Livett, 1971; Livett, 1973). Interestingly, an analogous situation has been described in several structures of the central nervous system (CNS), such as the cerebral cortex (Beaudet and Descarries, 1978; Descarries et al., 1977), amygdala (Farb et al., 2010) and hypothalamus (Marrotte and Raisman, 1974). In fact, the release of noradrenaline from axonal varicosi-ties is now regarded as a general feature of central noradrenergic neurones (Chiti and Teschemacher, 2007; Kasparov and Teschemacher, 2008). Although some of the noradrenergic bou-tons form one-to-one contacts with adjoining neurones, many of them are located in interstitial spaces without any identifiable syn-aptic contacts. The existence of non-synaptic boutons suggests that noradrenaline, apart from acting as synaptic transmitter, may also have a more diffuse hormone-like action in the brain (Beaudet and Descarries, 1978; Descarries and Mechawar, 2000). Interestingly, such a dual synaptic and non-synaptic action has also been demonstrated for the other monoamine neurotransmit-ters, dopamine and 5-hydroxytryptamine (serotonin) (Descarries and Mechawar, 2000) and for neuropeptide transmitters (Del Cid-Pellitero and GarzÓn, 2011). The terms ‘wiring transmission‘ and ‘volume transmission‘ have been used to refer to the two kinds of neurotransmission (Fuxe et al., 2010). A different terminology has been suggested by Nieuwenhuys (1985), referring to the synaptic

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 661

release of the neurotransmitter as ‘neurocrine secretion‘ and the non-synaptic release as ‘paracrine secretion‘.

Noradrenergic neurones may contain and release neurotrans-mitters other than noradrenaline (‘cotransmission‘: Burnstock, 2004; Campbell, 1987; Kupfermann, 1991). Cotransmitters in peripheral (sympathetic) noradrenergic neurones include adeno-sine triphosphate (ATP) and neuropetide Y (NPY) (Burnstock, 2004, 2009; Wier et al., 2009), and in central noradrenergic neu-rones ATP (Burnstock, 2004, 2009), NPY (Xu et al., 1998), gluta-mate (Fung et al., 1994a, 1994b), enkephalin (Fung et al., 1994a, 1994b) and the neuropeptide galanin (Le Maître et al., 2013; Xu et al., 1998). The cotransmitter may modulate the action of noradrenaline both at pre-synaptic sites (release) and post-synap-tic sites (neurotransmission) (Burnstock 2004, 2009; Herring and Paterson, 2009).

Central noradrenergic neurones make contacts not only with neurones but also with non-neuronal elements of the CNS, such as glia cells and blood vessels. Both types of macroglial cells (astro-cytes and oligodendrocytes) have been shown to possess intimate contacts with noradrenergic nerve fibers (Paspalas and Papadopoulos, 1996). Both α- and β-adrenoceptors have been identified on glia cells. The α-adrenoceptors belong mainly to the α1 subclass, whereas both β1 and β2 adrenoceptors are present in large densities on astrocytes (Hertz et al., 2004; Stone and Ariano, 1989). Recently β3-adrenoceptors have also been identified on astrocytes (Catus et al., 2011). The noradrenergic innervation of astrocytes has been implicated in the generalized metabolic response of the brain to stress via the activation of glycogenolysis, production of neurotrophic substances (e.g. nerve growth factor), mediation of glial-neuronal signaling (Farb et al., 2010; Stone and Ariano, 1989), and consolidation of memory (Hertz et al., 2004). Less information is available about the functional relevance of the noradrenergic innervation of oligodendrocytes. The major func-tion of oligodendroglia is the production of myelin: the myelin sheath of axons allows for the fast transmission of signals via ‘sal-tatory conduction‘ (Baumann and Pham-Dinh, 2001). While cen-tral noradrenergic axons, like the axons of peripheral sympathetic neurones, are unmyelinated (Kasparov and Teschemacher, 2008; Kiernan, 2005), the noradrenergic innervation of oligodendro-cytes may play an important role in the myelination of other path-ways (Davis et al., 2012).

Small blood vessels also receive an innervation from central noradrenergic neurones (Hartman et al., 1972; Paspalas and Papadopoulos, 1996; Raichle et al., 1975). This innervation has been implicated in the control of cerebral microcirculation: while the large extraparenchymal intracranial blood vessels receive their sympathetic noradrenergic innervation from the superior cervical ganglion, the small intraprenchymal vessels are under direct cen-tral noradrenergic control (Raichle et al., 1975). Both α- and β-adrenoceptors have been detected in cerebral microvessels (Kobayashi et al., 1982; Yokoo et al., 2000), and the α-adrenoceptors have been further characterized as belonging mainly to the α1A subtype (Yokoo et al., 2000). Although there are direct contacts between noradrenergic nerve terminals and cere-bral microvessels, in a number of instances the noradrenergic influence on the blood vessel seems to be mediated by glia cells, as indicated by the interposition of an astrocytic foot between the noradrenergic terminal and the vessel (Cohen et al., 1997; Paspalas and Papadopoulos, 1996).

As the peripheral and central neurones have common ori-gins, they share their genetic make-up and biochemical,

physiological and pharmacological properties. Figure 1 shows the pharmacologically sensitive sites of the noradrenergic neurone. In both peripheral and central noradrenergic neu-rones, noradrenaline is synthesized from the amino acid tyros-ine via the steps tyrosine → dopa → dopamine → noradrenaline, and stored in synaptic vesicles in the nerve terminal. The oper-ation of distinct transporter proteins in the vesicular and cel-lular membranes provides the mechanism for the accumulation and conservation of the transmitter within the synaptic vesi-cles and nerve terminals. Adrenoceptors stimulated by the released noradrenaline occur both post-synaptically and on the noradrenergic neurone itself (‘autoreceptors‘). Postsynaptic adrenoceptors in the periphery are located on smooth muscle or gland cells, and in the CNS on follower neurones and glia cells, and belong to any sub-class (α1, α2 or β) of adrenoceptor (Szabadi and Bradshaw, 1991; Szabadi et al., 1985). The auto-receptors are of the α2 sub-class and inhibitory in action: in a somato-dendritic location they inhibit neuronal firing (Huang et al., 2012), while located on the nerve terminal membrane (‘presynaptic receptors‘) they inhibit noradrenaline release (Starke, 2001). Noradrenergic neuronal activity is auto-regu-lated by the reuptake and re-storage of the released transmit-ter, and inhibition of noradrenaline release and cellular firing via the activation of inhibitory autoreceptors.

Interestingly, presynaptic release-modulating adrenoceptors occur not only on the axon terminals of noradrenergic neurones (autoreceptors), but also on the nerve terminals of other neurones (heteroreceptors) (Gilsbach and Hein, 2012; Starke, 2001). The presynaptic heteroreceptors, like the autoreceptors, belong to the α2 subclass and are inhibitory in nature. In the CNS, presynaptic heteroreceptors have been identified on the terminals of seroton-ergic (Millan et al., 2000; Scheibner et al., 2001), dopaminergic (Millan et al., 2000) and glutamatergic (Jiménez-Rivera et al., 2012; Shields et al., 2009) neurones. Furthermore, release-inhibit-ing adrenoceptors have been described on the nerve terminals of cholinergic neurones in the myenteric plexus (‘enteric nervous system‘) (LePard et al., 2004; Scheibner et al., 2002).

The three major sub-classes of adrenoceptor comprise a num-ber of subtypes that have been characterized pharmacologically on the basis of their affinities for selective ligands. Most of the receptor subtypes have been genetically defined and cloned. The α1-adrenoceptor sub-class comprises the subtypes α1A, α1B and α1D, and the α2-adrenoceptor sub-class consists of the subtypes α2A/D, α2B and α2C (Civantos Calzada and Aleixandre de Artiñano, 2001). The β-adrenoceptor sub-class comprises three subtypes: β1, β2 and β3 (Hieble and Ruffolo, 1995). All adrenoceptor subtypes have been identified in the CNS: α1-adrenoceptor subtypes (Day et al., 1997; Domyancic and Morilak, 1997; Shen et al., 2000; Szot et al., 2005; Williams and Morilak, 1997), α2-adrenoceptor subtypes (Flügge et al., 2003; Guyenet et al., 1994; Owesson et al., 2003; Scheinin et al., 1994; Tavares et al., 1996), and β-adrenoceptor subtypes (Hillman et al., 2005; Joyce et al., 1992; Paschalis et al., 2009; Summers et al., 1995; Ursino et al., 2009). Presynaptic adrenoceptors, both autoreceptors and heterorecep-tors, belong to the α2A/D subtype (Gilsbach and Hein, 2012; Starke, 2001).

3. Noradrenergic nucleiNoradrenergic neurones occur in clusters both in the periphery (sympathetic ganglia) and in the CNS (noradrenergic nuclei). A

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

662 Journal of Psychopharmacology 27(8)

general feature of the central noradrenergic system is that the cell bodies are located in a relatively small number of distinct nuclei in the lower brainstem (pons and medulla oblongata), and the axons project to most parts of the brain and the spinal cord. There is a diffuse projection from the brainstem nuclei: many axon collater-als may arise from one neurone which may reach several remote areas of the CNS (e.g. cerebral cortex, limbic areas, hypothala-mus, cerebellum, spinal cord). However, in parallel with the dif-fuse noradrenergic projection, more targeted outputs to distinct destinations have also been identified (Nieuwenhuys, 1985; Papadopoulos and Parnavelas, 1991).

Seven noradrenergic nuclei can be distinguished, and they are labeled A1–A7 (Dahlström and Fuxe, 1964). Although these nuclei have originally been described in rats, most of them have also been identified in primates (Felten and Sladek, 1983; Jacobowitz and MacLean, 1978), including man (Bogerts, 1981). This suggests that the central noradrenergic system is a develop-mentally ancient system of the brain.

The noradrenergic nuclei can be divided into three groups: caudal (or medullary) group, central (or medullo-pontine) group, and rostral (or pontine) group. The rostral group is the most important in man: it consists of only one nucleus, area A6 (LC). The LC is the largest noradrenergic nucleus: it contains almost 50% of all the noradrenergic neurones. The LC is located in the dorsal tegmentum of the rostral area of the pons: it appears as a

bluish-grey (‘coeruleus‘) patch on the floor of the fourth cere-bral ventricle (Figure 2). The color of the LC is due to the pres-ence of the pigment neuromelanin in LC neurones. The presence of neuromelanin has led to the development of the technique of neuromelanin magnetic resonance imaging which has enabled the visualization of the LC in the human brain in vivo (Shibata et al., 2006) (Figure 2).

The fiber projections arising from the noradrenergic nuclei can be divided into two major groups: fibers arising from the caudal and central nuclear groups (A1, A2, A5 and A7) and fibers originating from the LC (A6). Fibres originating from areas A1, A2, A5 and A7 give rise to both ascending and descending fiber systems. The ascending fiber system forms the ventral noradrenergic bundle which innervates the mid-brain reticular formation, the entire hypothalamus and parts of the limbic cortex. The descending fibers form two bulbos-pinal bundles which terminate in the ventral and the interme-diate nuclear columns of the spinal cord. It should be noted that many of the targets of the ventral noradrenergic pathway also receive inputs from the LC. Thus, for example, while a major noradrenergic input to the hypothalamus derives from the A1, A2, A5 and A7 nuclei (ventral noradrenergic bundle), the LC also contributes to the noradrenergic innervation of the hypothalamus via the dorsal noradrenergic bundle (Nieuwenhuys, 1985).

Figure 1. Pharmacologically sensitive sites of the noradrenergic neurone. (Left) Adrenoceptors. Noradrenaline released from the axon terminals of the neurone can activate postsynaptic adrenoceptors on the target cells. Adrenoceptors are of three classes: α1 (excitatory); α2 (inhibitory); β (excitatory or inhibitory). There are also adrenoceptors on the noradrenergic neurone itself (autoreceptors): these are of the α2 subtype, and mediate inhibition. The autoreceptors are stimulated by noradrenaline released from the noradrenergic neurone. Terminal autoreceptors are located on the nerve terminals (‘presynaptic receptors‘) and inhibit the release of noradrenaline (‘release-modulating receptors’): somatodendritic autoreceptors are located on the cell body or the dendrites, and inhibit neuronal firing. Somatodendritic autoreceptors are stimulated by noradrenaline released either from the dendrites (‘dendritic release‘) or from recurrent nerve terminals. The stimulation of autoreceptors dampens the activity of the neurone (‘autoregulation‘). (Right) Noradrenergic synapse. Noradrenaline released from the nerve terminal stimulates postsynaptic receptors (1); it is eliminated from the synaptic gap by reuptake into the nerve terminal (2), and then can be taken up into the synaptic vesicle via vesicular uptake and re-stored in the vesicle (3).

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 663

The fiber projections from the LC give rise to three path-ways: the ascending pathway (dorsal noradrenergic bundle), the cerebellar pathway and the descending pathway. The ascending pathway innervates structures in the midbrain (periaqueductal grey substance, nucleus raphe dorsalis, colliculi), thalamus, lim-bic system (amygdala, hippocampus, cingulate and parahip-pocampal gyri), and all areas of neocortex. The cerebellar pathway projects to the cerebellar nuclei and cerebellar cortex via the superior cerebellar peduncle. The descending pathway sends collaterals to motor nuclei in the lower brainstem (dorsal nucleus of the vagus, inferior olivary complex), and then descends to the spinal cord (coeruleo-spinal pathway), innervat-ing spinal neurones in all three nuclear columns. The functions of the noradrenergic pathways are discussed in detail with their targets in Section 4.

Although the organization of the noradrenergic nuclei and their fiber projections show great similarity between different mammalian species, there are also some differences. In pri-mates, nucleus A3 is absent (Nieuwenhuys, 1985) and while the medullary tegmental nuclei (A1 and A2) are very similar to those in rodents, the pontine nuclei (A4, A5, A6, A7) con-tain a much larger number of noradrenergic neurones than in

rodents (Schofield and Everitt, 1981). In carnivores (e.g. cat: Poitras and Parent, 1978; dog: Ishikawa et al., 1975), the pon-tine nuclei have large noradrenergic cell numbers like those in primates, however, the neurones are less aggregated into distinct nuclei and have a more diffusely distributed localiza-tion. In ungulates (e.g. sheep: Tillet and Thibault, 1989), the noradrenergic neurones are more loosely clustered than in rodents and primates, and the number of neurones is lower in the medullary noradrenergic nuclei (Tillet and Kitahama, 1998). For a detailed discussion of the species differences in the organization of central noradrenergic nuclei and the dis-tribution of their fiber projections, see Schofield and Everitt (1981). Although the anatomical differences between species are significant, they are relatively minor compared to differ-ences between the functional markers of central noradrener-gic activity. Thus major species differences in the distributions of both α- and β-adrenoceptors (Booze et al., 1989; Palacios et al., 1987) and the localization of the peptide co-transmitter galanin (Le Maître et al., 2013) have been reported. These observations indicate that caution is required when results obtained in rodents are extrapolated to other species, includ-ing man.

Figure 2. Location of the locus coeruleus (LC) in the brainstem. (Top left) Sagittal section through the brainstem showing the level of the cross-section displayed on the right. (Top right) Horizontal section through the upper pons showing the fourth cerebral ventricle. The loci coerulei are located in the lateral corners of the ventricle (shown in red). Reproduced, with permission, from Warner (2001) (copyright: Elsevier).(Bottom) Nuclear magnetic resonance image showing a cross-section of the upper pons displaying the loci coerulei. The picture was taken in a healthy human subject, with the modification of the method of Sasaki et al. (2006), on a 3-tesla scanner to obtain a neuromelanin signal to identify the loci coerulei. The loci coerulei are shown by the small white areas in the bottom corners of the fourth cerebral ventricle, indicated by yellow arrows. By courtesy of D Auer, Queen‘s Medical Centre, Nottingham. Reproduced from Samuels ER and Szabadi E (2008b) Functional neuroanatomy of the norad-renergic locus coeruleus: Its roles in the regulation of arousal and autonomic function. Part II: Physiological and pharmacological manipulations and pathological alterations of locus coeruleus activity in humans. Curr Neuropharmacol 6: 254–285 with permission from Bentham Science Publishers.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

664 Journal of Psychopharmacology 27(8)

4. Efferents of the LCFigure 3 shows a schematic overview of LC efferents.

4.1 Telencephalon

The telencephalon is the most anterior part of the forebrain, com-prising the cerebral cortex and the basal ganglia. Other parts of the forebrain include the limbic system and the diencephalon.

4.1.1 Neocortex. The neocortex receives a rich noradrenergic innervation which derives entirely from the LC. Indeed, there is a close correlation between LC activity and noradrenaline release in the neocortex (Berridge and Abercrombie, 1999).

The LC projects (coeruleo-cortical pathway), via the dorsal noradrenergic bundle, to all areas of the neocortex (Gatter and Powell, 1977; Jones and Yang, 1985; Séguéla et al., 1990). The extensive nature of the noradrenergic innervation of the neocortex suggests that the projection from the LC exerts a diffuse influence on this structure (Nieuwenhuys, 1985). Although there is anatomi-cal evidence consistent with a diffuse pattern of noradrenergic innervation (Loughlin et al., 1982; Nagai et al., 1981), a large degree of topographic specificity has also been found (Foote et al., 1983; Levitt et al., 1984; Morrison et al., 1979, 1982; Papadopoulos and Parnavelas, 1991). This topographic specificity is illustrated by the recent observation that LC neurones that receive afferents from the orexinergic neurones of the lateral hypothalamic area project to the medial prefrontal cortex (Del Cid-Pellitero and Garzόn, 2011).

Figure 3. Efferents of the locus coeruleus (LC). Arrows indicate projections. Red arrowhead: excitatory; blue arrowhead: inhibitory. Coloring of target area indicates function (for code see figure).A: amygdala; BF: basal forebrain; Cllm: cerebellum; CR: caudal raphe; DR: dorsal raphe; Hc: hippocampus; Ht: hypothalamus; IL: intralaminar nuclei; LH: lateral hypo-thalamic area; PPT/LDT; pedunculopontine/laterodorsal tegmental nuclei; PVN: paraventricular nucleus; RVLM: rostroventrolateral medulla; SS: somatosensory thalamus; symp: sympathetic preganglionic neurones; Tinf: tuberoinfundibular area; Th thalamus; VLPO: ventrolateral preoptic area.In the brainstem, encircled Arabic numbers indicate functional nuclear groups: (1), motor; (2), sensory; (3), premotor sympathetic; (4), preganglionic parasympathetic; (5), sleep/wakefulness-modulating. Roman numbers correspond to cranial nerves: III: oculomotor; V: trigeminal; VII: facial; VIII: auditory; IX: glossopharyngeal; X: vagus. The LC exerts an excitatory influence on the neocortex, hippocampus, amygdala and cerebellum, on wakefulness-promoting neurones in the BF, thalamus (IL) and brainstem (DR, PPT/LDT), on motor neurones in the brainstem and spinal cord, on sensory neurones in the thalamus (SS) and brainstem, on premotor sympathetic neu-rones in the hypothalamus (PVN) and brainstem (CR) and on neuroendocrine neurones in the hypothalamus (PVN and Tinf), and on preganglionic sympathetic neurones in the spinal cord. It should be noted, however, that the noradrenergic innervation of the preganglionic sympathetic neurones derives mainly from the A5/A7 pontine nuclei. On the other hand, the LC inhibits sleep-promoting neurones in the BF, hypothalamus (VLPO) and brainstem (PPT/LDT), some wakefulness-promoting neurones in the hypothalamus (LH), some premotor sympathetic neurones in the brainstem (RVLM), preganglionic parasympathetic neurones in the brainstem, some trigeminal (not shown, see Section 5.4.3) and spinal sensory neurones.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 665

The LC exerts an excitatory influence on the cerebral cortex. This is largely due to the stimulation of excitatory α1-adrenoceptors that are expressed in high concentrations through-out the cortex (Domyancic and Morilak, 1997; Jones et al., 1985; Papay et al., 2006; Pieribone et al., 1994). All three subtypes of α1-adrenoceptor have been identified in the neocortex: α1A (Day et al., 1997; Domyancic and Morilak, 1997; Pieribone et al., 1994), α1B (Acosta-Martinez et al., 1999; Day et al., 1997; Pieribone et al., 1994) and α1D (Day et al., 1997; Pieribone et al., 1994). Inhibitory α2-adrenoceptors have also been identified, however, in smaller numbers than the α1-adrenoceptors and in a more selective distribution (Pascual et al., 1992; Talley at al., 1996). The cortical α2-adrenoceptors belong mainly to the α2A subtype (Blake et al., 1998; Scheinin et al., 1994). It has been proposed that these receptors may be located on inhibitory interneurones where their stimulation would disinhibit cortical neurones leading to cortical excitation (Andrews and Lavin, 2006). β-Adrenoceptors are also detectable in the neocortex (Wanaka et al., 1989): all three subtypes have been identified (Chen et al., 2012; Joyce et al., 1992; Paschalis et al., 2009; Rodriguez et al., 1995; Van Warde et al., 1997).

There is a close relationship between LC activity and overall cortical activity as shown by the electroencephalogram (EEG): activation of the LC leads to the activation of the cerebral cortex (Berridge and Foote, 1991), whereas inactivation of the LC results in a reduction in cortical EEG activity (Berridge et al., 1993; Danysz et al., 1989). Cortical activation by the LC under-lies the role of the LC as a major wakefulness-promoting nucleus (see Section 6.1). Indeed, there is a close parallelism between LC activity and the level of arousal (Aston-Jones and Bloom, 1981; Aston-Jones et al., 1991a; Foote et al., 1991; Rajkowski et al., 1994).

4.1.2 Basal forebrain. The BF comprises the medial septal area, medial preoptic area and substantia innominata. All three areas receive a noradrenergic innervation from the LC (España and Berridge, 2006). The BF contains both cholinergic excitatory and γ-aminobutyric acid (GABA)ergic inhibitory neurones (Gritti et al., 1993). Both types of neurone project to the neocortex and play a role in the modulation of arousal (Jones, 2004; Modirrousta et al., 2004): the cholinergic neurones are active during wakeful-ness and quiescent during sleep (Manns et al., 2000a) whereas the reverse pattern of activity applies to the GABAergic neurones (Manns et al., 2000b). The noradrenergic projection from the LC modulates the activity of both types of neurone (Figure 4). Nor-adrenaline released from the noradrenergic terminals activates the wakefulness-promoting cholinergic neurones (Fort et al., 1995; Jones, 2004), via the stimulation of excitatory α1- and β1-adrenoceptors (Berridge et al., 2003) whereas it suppresses the activity of sleep-promoting GABAergic neurones, via the stimula-tion of inhibitory α2-adrenoceptors (Manns et al., 2003). Overall, the LC exerts a wakefulness-promoting effect via the BF. Thus, the indirect activation of the neocortex via the BF enhances the wakefulness-promoting effect of the LC resulting from cortical activation by the coeruleo-cortical pathway (see above).

4.2 Limbic system

4.2.1 Amygdala. The LC sends a dense noradrenergic projec-tion to the amygdala, mainly to the central and basal nuclei (Fal-lon et al., 1978; Jones and Moore, 1977). The adrenoceptors expressed in the amygdala are mainly of the α1 sub-class (Jones et al., 1985; Papay et al., 2006; Pieribone et al., 1994), although a smaller population of α2-adrenoceptors has also been identified (Glass et al., 2002; Pascual et al., 1992). The α1- adrenoceptors

Figure 4. Dual influence of the basal forebrain (BF) on cortical arousal. Circles represent neurones and arrows represent connections (red, excitatory; blue, inhibitory). Neurotransmitters: ACh: acetylcholine; GABA: γ-aminobutyric acid; NA: noradrenaline. Cholinergic neurones of the BF stimulate, and GABAergic neurones inhibit, cortical activity. The locus coeruleus (LC) modulates BF activity by a dual action: it stimulates the excitatory cholinergic neurones via α1-adrenoceptors, and inhibits the inhibitory GABAergic neurones via α2-adrenoceptors. Overall, the LC exerts a powerful wakefulness-promoting influence via the BF (see Section 4.1.2 for details).

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

666 Journal of Psychopharmacology 27(8)

belong mainly to the α1A subtype (Day et al., 1997), and the α2- adrenoceptors to the α2A subtype (Glass et al., 2002; Scheinin et al., 1994). β-Adrenoceptors are also present in the amygdala, both in the central and basolateral nuclei (Wanaka et al., 1989): they have been identified as belonging to the β1 (Ghiasvand et al., 2011; Paschalis et al., 2009; Qu et al., 2008; Silberman et al., 2012) and the β2 (Qu et al., 2008) subtypes. Specific functions have been attributed to β-adrenoceptors in different amygdaloid nuclei: β-adrenoceptors in the lateral nucleus have been impli-cated in fear-conditioning (Farb et al., 2010), in the basolateral nucleus in anxiety (Silberman et al., 2012), in consolidation of fear memory (Qu et al., 2008) and in the central nucleus in state-dependent memory (Ghiasvand et al., 2011). Both excitatory and inhibitory neuronal responses to locally applied noradrenaline and to LC stimulation have been observed in the basolateral nucleus, the excitatory responses being mediated via β-adrenoceptors and the inhibitory responses via α2-adrenoceptors (Buffalari and Grace, 2007).

The amygdala plays a key role in mediating fear and anxiety responses to threatening environmental stimuli (Damasio, 1998; LeDoux, 1998). The stimulation of the LC, electrically or by the administration of an α2-adrenoceptor antagonist (e.g. yohimbine), leads to an increase in anxiety (McDougle et al., 1995; Redmond and Huang, 1979), consistent with the activation of the amygdala which suggests that the overall noradrenergic influence on the amygdala is likely to be excitatory. The noradrenergic projection to the amygdala has also been implicated in the formation and retrieval of emotional memories (Chen and Sara, 2007; Sterpenich et al., 2006).

4.2.2 Hippocampus. The LC also sends a projection to the hip-pocampus: in fact, the LC is the sole source of noradrenaline in this structure (Fu et al., 1999; Loughlin et al., 1986b). Both α1- (Jones et al., 1985; Pieribone et al., 1994) and α2-adrenoceptors (Pascual et al., 1992; Scheinin et al., 1994) have been identified in the hippocampus. There is a species difference in the densities of hippocampal α1-adrenoceptors between primates and rodents: these receptors are much more numerous in primates than in rodents (Palacios et al., 1987). The α1-adrenoceptors belong mainly to the α1A subtype (Day et al., 1997; O‘Malley et al., 1998; Pieribone et al., 1994; Szot et al., 2005), although α1D adrenocep-tors are also present (Williams et al., 1997). The α2-adrenoceptors belong mainly to the α2A subtype (Scheinin et al., 1994). β-Adrenoceptors are also present in the hippocampus (Wanaka et al., 1989): both β1 and β2 subtypes have been identified (Guo and Li, 2007; Hillman et al., 2005; Joyce et al., 1992; Paschalis et al., 2009).There is also evidence of the presence of β3 adreno-ceptors in the hippocampus (Summers et al., 1995). The LC pro-jection to the hippocampus has been implicated both in the formation (Sullivan et al., 1994) and retrieval (Sara and Devauges, 1988) of memories.

4.3 Diencephalon

The diencephalon consists of the thalamus and hypothalamus.

4.3.1 Thalamus. The thalamus receives a profuse noradrener-gic projection from the LC (Ishikawa and Tanaka, 1977; Jones and Yang, 1985). Excitatory α1-adrenoceptors are abundant in

the thalamus (Jones et al., 1985): they belong to the α1B subtype (Day et al., 1997; Pieribone et al., 1994). A smaller population of α2-adrenoceptors is also present in the thalamus: these belong to the α2B subtype (Scheinin et al., 1994). β1-Adrenoceptors have been identified in the medial habenular nucleus of the thalamus, but most other areas of the thalamus are devoid of this adreno-ceptor sub-class (Paschalis et al., 2009). The noradrenergic pro-jection to the thalamus has been implicated in the promotion of wakefulness (McBride and Sutin, 1976; McCormick et al., 1991) and the modulation of pain (Westlund et al., 1991; Zhang et al., 1998a).

4.3.2 Hypothalamus. Although the noradrenergic innervation of the hypothalamus derives largely from the A1/A5 nuclei via the ventral noradrenergic bundle, the LC also contributes to it (Cun-ningham and Sawchenko, 1988; Niewenhuys, 1985).

Two arousal-modulating hypothalamic nuclei, the ventrolat-eral preoptic area (VLPO) and the lateral hypothalamic/periforni-cal area (LH/PF) are under noradrenergic influence.

4.3.2.1 VLPO. The VLPO, the major sleep-promoting nucleus, receives a noradrenergic input from the LC (Chou et al., 2002; Osaka and Matsumura, 1994). Noradrenaline, released from LC nerve terminals during wakefulness, inhibits GABAergic VLPO neurones via the stimulation of α2-adrenoceptors (Gallo-pin et al., 2000; Modirrousta et al., 2004; Osaka and Matsumura, 1994), and thereby promotes wakefulness (Nelson et al., 2002). The α2-adrenoceptors in this nucleus belong to the α2A subtype (Scheinin et al., 1994).

4.3.2.2 LH/PF. The LC also projects to the wakefulness-promoting orexinergic neurones in the LH/PF (Baldo et al., 2003; Yamanaka et al., 2003). There is only a sparse presence of α1-adrenoceptors in the lateral hypothalamus, mainly in the form of α1B-adrenoceptors (Day et al., 1997). On the other hand, α2-adrenoceptors are more abundant (α2A subtype: Scheinin et al., 1994). The noradrenergic projection to this area exerts an inhibitory influence on the orexinergic neurones (Li and van den Pol, 2005; Yamanaka et al., 2003), probably via the stimulation of inhibitory α2A-adrenoceptors. Interestingly, these α2-adrenoceptors seem to be subject to regulation by sleep-depriva-tion: it has been reported that the α2-adrenoceptor agonist clonidine inhibits orexinergic neurones only in sleep-deprived animals, whereas it is ineffective in non-sleep deprived controls (Uschakov et al., 2011). As the orexinergic neurones send a powerful excitatory projection to the LC (see Section 5.3.2), the inhibitory output from the LC to the LH/PF constitutes a nega-tive feedback loop which may serve to protect the LC from over-excitation during arousal (Samuels and Szabadi, 2008a). For details of the role of the LC in the regulation of sleep and arousal, see Section 6.1.

4.3.2.3 Paraventricular nucleus. The LC sends a projec-tion to the paraventricular nucleus (PVN) of the hypothalamus (Jones and Moore, 1977; Jones and Yang, 1985). The PVN is a major premotor autonomic nucleus that modulates the activity of both sympathetic and parasympathetic preganglionic neurones (i.e. ‘vegetative motor neurones‘) (Buijs et al., 2003; Swanson and Sawchenko, 1980, 1983). The sympathetic preganglionic neurones are located in the intermedio-lateral nuclear column (IML) of the spinal cord, whereas the parasympathetic pregangli-onic neurones are situated in the medulla oblongata. The PVN projects to both these sites (Holstege, 1987; Hosoya et al., 1991;

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 667

Papay et al., 2006; Pyner, 2009; Ranson et al., 1998; Saper et al., 1976; Sawchenko and Swanson, 1982). Both α1- (Chong et al., 2004; Day et al., 1997) and α2-adrenoceptors (Chong et al., 2004; Li et al., 2005) are expressed in the PVN. The α1-adrenoceptors belong to the α1A (Chong et al., 2004; Day et al., 1997) and the α1B subtypes (Williams and Morilak, 1997) and the α2-adrenoceptors to the α2A subtype (Chong et al., 2004; Scheinin et al., 1994). β-Adrenoceptors are also present in this nucleus (Boundy and Cincotta, 2000; Takano et al., 1989; Wanaka et al., 1989), mainly in the form of β1-adrenoceptors (Paschalis et al., 2009). The acti-vation of both α-adrenoceptor subtypes leads to excitation of PVN neurones: α1-adrenoceptor activation directly, and α2-adrenoceptor activation indirectly, by switching off the activity of GABAergic inhibitory interneurones on which these receptors are located (Li et al., 2005). Apart from the premotor autonomic neu-rones, there is also a population of smaller (‘parvocellular‘) neu-rones in the PVN which have neurosecretory (neuroendocrine) functions (Cunningham and Sawchenko, 1988; Daftary et al., 2000; Swanson and Sawchenko, 1980). These neurones, that secrete trope hormones regulating pituitary activity, are also under noradrenergic influence (Sands and Morilak, 1999). Noradrena-line excites the majority of parvocellular neurones via α1-adrenoceptors, but inhibits a subpopulation via β-adrenoceptors (Daftary et al., 2000). The noradrenergic projection to the PVN has been implicated in mediating autonomic and endocrine responses to stress (Bienkowski and Rinaman, 2008; Herman et al., 2003; Li et al., 1996).

4.3.2.4 Arcuate nucleus/tuberinfundibular area. The arcuate nucleus of the tuberinfundibular area of the hypothala-mus also receives a noradrenergic input both from the LC (Jones and Moore, 1977; Pickel et al., 1974), and the A1 noradrener-gic nucleus in the medulla (Palkovits et al., 1980). The arcuate nucleus, like the PVN, is a major neuroendocrine nucleus which, through its manifold connections, is involved in the integration of endocrine, emotional, sensory, vegetative homeostatic and autonomic functions (Chronwall, 1985). Both excitatory α1- and β-adrenoceptors (Acosta-Martinez et al., 1999; Kang et al., 2000) and inhibitory α2-adrenoceptors (Young and Kuhar, 1980) have been identified in the arcuate nucleus. The α1-adrenoceptors belong mainly to the α1A subtype (Day et al., 1997), the α2-adrenoceptors to the α2A subtype (Scheinin et al., 1994) and the β-adrenoceptors to the β1 subtype (Paschalis et al., 2009). The secretion of growth hormone (GH) by the pituitary is under con-trol by growth-hormone-releasing hormone (GHRH) that stim-ulates GH secretion and somatostatin that inhibits it (Guistina and Veldhuis, 1998). GHRH and somatostatin are secreted by neurosecretory peptidergic arcuate neurones into the portal cir-culation. The secretion of GH is modulated by the noradrenergic system via α2-adrenoceptors. Although it is well documented that the stimulation of central α2-adrenoceptors (e.g. by the admin-istration of the α2-adrenoceptor agonist clonidine) leads to an increase in GH secretion, the location of these receptors remains controversial, both GHRH- and somatostatin-secreting arcuate neurones having been implicated (Giustina and Veldhuis, 1998). Dopaminergic neurones in the arcuate nucleus (‘tuberoinfundib-ular dopamine (TIDA) neurones‘) are responsible for the tonic inhibition of the release of prolactin by the pituitary (Ben-Jon-athan and Hnasko, 2001; Lyons et al., 2012). The noradrenergic system is likely to be involved in the regulation of the activity of TIDA neurones, since it has been shown that these neurones

receive a noradrenergic input (Hrabovsky and Liposits, 1994). Furthermore, it has been reported that the wakefulness-promot-ing drug modafinil, believed to enhance noradrenergic LC activ-ity (Hou et al., 2005), reduces prolactin secretion (Samuels et al., 2006) which is consistent with the stimulation of TIDA neurones via α1-adrenoceptors (Kang et al., 2000).

4.4 Brainstem

The LC projects, together with other noradrenergic nuclei, to a large number of neuronal groups in the brainstem. These include autonomic (sympathetic premotor and parasympathetic pregan-glionic) nuclei, somatic (motor and sensory) nuclei, and also nuclei involved in the regulation of sleep and arousal. The para-sympathetic preganglionic nuclei, together with the somatic motor and sensory nuclei, are usually described under the head-ing ‘cranial nerves‘.

4.4.1. Sympathetic premotor nuclei. Two important neuronal groups in the medulla have higher order (i.e. premotor) sympa-thetic functions: the rostroventrolateral medulla (RVLM) and the caudal raphe (CR) nuclei.

4.4.1.1 RVLM. Glutamatergic neurones of the RVLM pro-ject to preganglionic sympathetic neurones in the IML of the spi-nal cord (Barman and Gebber, 1985; Zagon and Smith, 1993). RVLM neurones possess an intrinsic pacemaker activity, and play a role in the maintenance of blood pressure and heart rate (Damp-ney, 1994; Sun, 1995). In general, the RVLM has a stimulatory effect on blood pressure (Kapoor et al., 1992) and also contributes to the efferent branch of the baroreflex (Dampney, 1994). The LC projects to the RVLM (Van Bockstaele et al., 1989) (‘coeruleo-vasomotor pathway‘), where noradrenaline exerts an inhibitory effect via the stimulation of α2–adrenoceptors (Head et al., 1998; Tavares at al., 1996). However, the stimulation of the LC is asso-ciated with moderate increases in heart rate and blood pressure (Drolet and Gauthier, 1985; Gurtu et al., 1984), suggesting that the vasodepressor effect of the LC, arising from the inhibition of the RVLM, may be superseded by a vasopressor effect. The likely substrate for this vasopressor effect is the stimulation of excita-tory α1-adrenoceptors on preganglionic sympathetic neurones via a direct output from the LC to the spinal cord (‘coeruleo-spinal pathway‘, see below). For a detailed description of the role of the LC in cardiovascular regulation, see Section 6.2.

4.4.1.2 CR nuclei. The nuclei (raphe magnus, obscurus and pallidus) of the CR are involved in modulating sympathetic functions via serotonergic outputs to the IML of the spinal cord (Allen and Cechetto, 1994; Loewy, 1981). The LC projects to the CR (Hermann et al., 1997). The adrenoceptors identified in the CR are mainly of the α1 subtype (Domyancic and Morilak, 1997), although α2-adrenoceptors have also been detected (Guy-enet et al., 1994). The CR has been implicated in mediating thermoregulatory functions (Nakamura et al., 2004, 2005): cold exposure excites CR neurones (Rathner et al., 2001). The noradr-energic stimulation of CR neurones by the LC would be expected to lead to sympathetic excitation resulting in an increase in body temperature. A subgroup of CR neurones projects to the dorsal column of the spinal cord (Kwiat and Basbaum, 1992; Rivot et al., 1982), and is involved in the suppression of nociception (Furst, 1999; Millan, 2002). The noradrenergic projection to the

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

668 Journal of Psychopharmacology 27(8)

antinociceptive CR neurones has been shown to modulate sero-tonergic antinociception (Bie et al., 2003).

4.4.2 Parasympathetic preganglionic nuclei. These nuclei include the Edinger-Westphal nucleus (EWN) (cranial nerve III), the salivatory nuclei (cranial nerves VII and IX), and the vagal nuclei (cranial nerve X). The preganglionic neurones are choliner-gic, and project to the cholinergic neurones of the parasympa-thetic ganglia. They receive a noradrenergic input from the LC which exerts an inhibitory action on them via the stimulation of α2-adrenoceptors.

4.4.2.1 EWN. This nucleus is part of the oculomotor nuclear complex of the midbrain, situated in the ventral periaqueductal grey on the dorsal tip of the somatomotor nucleus of the third cra-nial nerve. EWN neurones project to postganglionic neurones in the ciliary ganglion (CG), which in turn innervate smooth muscle fibers in the sphincter pupillae muscle of the iris and the ciliary body. EWN neurones are light-sensitive: light stimulates mel-anopsin-containing intrinsically photosensitive retinal ganglion cells (ipRGCs), which project to the EWN via the olivary pretec-tal nucleus (OPN) (Gooley et al., 2003; Kawasaki and Kardon, 2007). Light evokes pupil constriction via the ‘retina → OPN → EWN → CG → sphincter pupillae muscle‘ pathway (‘pupillary light reflex‘) (Thompson, 1996). The EWN receives a noradren-ergic input that may originate from both both the A5/A7 nuclei (Levitt and Moore, 1979) and the LC (Breen et al., 1983). The noradrenergic innervation inhibits the activity of the EWN neu-rones via the stimulation of α2-adrenoceptors (Koss, 1986, Koss et al., 1984). Indeed, there is evidence that LC activation (e.g. by anxiety) leads to the inhibition of the pupillary light reflex (Samu-els and Szabadi, 2008b).

4.4.2.2 Salivatory nuclei. The two salivatory nuclei are located in the medulla: the superior salivatory nucleus in a dor-solateral, and the inferior salivatory nucleus in a ventrolateral position (Nicholson and Severin, 1981). The superior salivatory nucleus projects, through the facial nerve (cranial nerve VII), to two parasympathetic ganglia: the submandibular ganglion which innervates the submandibular and sublingual salivary glands (Chibuzo and Cummings, 1980); and the pterygopalatine gan-glion which innervates the lacrimal gland (Tóth et al., 1999). The inferior salivatory nucleus projects, through the glossopharyngeal nerve (cranial nerve IX), to the otic ganglion, from which post-ganglionic fibers innervate the parotid gland and the lingual (von Ebner) salivary glands (Fukami and Bradley, 2005; Kim et al., 2004a). Noradrenergic nerve terminals have been detected in the salivatory nuclei (Nemoto et al., 1995). The noradrenergic inner-vation of the preganglionic salivatory neurones has been traced back to the A5 pontine noradrenergic nucleus (Spencer et al., 1990) and the LC (Jansen et al., 1992). Inhibitory α2-adrenoceptors have been identified on salivatory neurones (Lung, 1994). It has been suggested that salivation is under tonic inhibitory noradrenergic control (Takakura et al., 2003).

4.4.2.3 Vagal nuclei. There are two vagal nuclei that con-tain preganglionic parasympathetic neurones: the dorsal motor nucleus of the vagus (DMV) and the nucleus ambiguous (Loewy and Spyer, 1990; Nosaka et al., 1982). These vagal nuclei are responsible for the parasympathetic nerve supply of the heart, respiratory tracts and lungs, nasopharynx, esophagus, gastroin-testinal tract, liver and pancreas. The parasympathetic ganglia to which the vagal nuclei project are located on the surface of the

organs supplied or embedded in their tissue. Both nuclei play an important role in cardiovascular regulation, however, the nucleus ambiguous plays the major role (Dampney, 1994). The integrity of the nucleus ambiguous is critical for the heart rate response to baroreceptor stimulation (Cheng et al., 2004). Injection of glutamate into the DMV (Chen and Chai, 2002) or the nucleus ambiguous (Machado and Brody, 1990) has been reported to induce bradycardia, consistent with a cardioinhibitory effect. The direct cardioinhibitory effect of the nucleus ambiguous via the vagus nerve is likely to be augmented be an inhibitory projection from the nucleus to the RVLM (McKitrick and Calaresu, 1996), a major sympathetic premotor nucleus involved in cardiovascular regulation (see above). The A5 noradrenergic nucleus projects to the DMV (Loewy et al., 1979), and the LC, via the coeruleo-vagal pathway, projects to both the DMV (Ter Horst et al., 1991; Westlund and Coulter, 1980) and the nucleus ambiguous (Jones and Yang, 1985; Westlund and Coulter, 1980). The central noradrenergic system exerts an inhibitory influence on the activ-ity of vagal parasympathetic preganglionic neurones: inhibitory α2-adrenoceptors have been identified in the DMV (Robertson and Leslie, 1985). The α2-adrenoceptors belong to the α2A sub-type (Flügge et al., 2003; Scheinin et al., 1994). Noradrenaline inhibits DMV neurones by stimulating these receptors (Fukuda et al., 1987; Martinez-Peña y Valenzuela et al., 2004). Apart from α2-adrenoceptors on DMV neurones, both α1A-adrenoceptors (Day et al., 1997) and β1-adrenoceptors (Bateman et al., 2012; Paschalis et al., 2009) have been detected in the nucleus ambigu-ous. For details of the role of the interplay between the LC and the vagal nuclei in cardiac control, see Section 6.2.

4.4.3 Motor nuclei. The LC projects to motoneurones both in the brainstem and spinal cord, exerting a facilitatory effect on these neurones via the activation of α1-adrenoceptors.

4.4.3.1 Oculomotor nucleus. This nucleus, belonging to the system of cranial nerve III, is responsible for the innervation of the majority of the external muscles of the eye, together with the levator palpebrae muscle. The nucleus has been shown to be supplied by noradrenergic fibers (Carpenter et al., 1992). α1-Adrenoceptors have been identified in the nucleus, with α1A being the dominant subtype (Day et al., 1997; Domyanic and Morilak, 1997; Pieribone et al., 1994).

4.4.3.2 Trigeminal motor nucleus. The trigeminal motor nucleus, belonging to the system of cranial nerve V, is responsi-ble for innervating the muscles involved in chewing. This nucleus receives its noradrenergic innervation mainly form the A1/A7 noradrenergic nuclei, with only a sparse contribution from the LC (Lyons and Grzanna, 1988). α1-Adrenoceptors are present in high density in this nucleus, with α1A being the most prevalent subtype (Day et al., 1997; Pieribone et al., 1994). Noradrenaline exerts a facilitatory influence on the activity of trigeminal motoneurones (Shao and Sutin, 1991).

4.4.3.3 Facial nucleus. The facial nucleus, belong-ing to the system of cranial nerve VII, is responsible for inner-vating the muscles of the face. It receives a rich noradrenergic input from the LC (Jones and Yang, 1985; McBride and Sutin, 1976). α1-Adrenoceptors are present in the facial nucleus, with α1A-adrenoceptors being the dominant subtype (Day et al., 1997; Domyanic and Morilak, 1997; Pieribone et al., 1994). Noradren-aline applied to facial motoneurones increases their activity (Larkman and Kelly, 1992; Rasmussen and Aghajanian, 1990;

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 669

VanderMaelen and Aghajanian, 1980). The excitatory projection from the LC to the facial motoneurones is instrumental in main-taining the tone of facial muscles.

4.4.3.4 Hypoglossal nucleus. This nucleus, belonging to the system of cranial nerve XII, is responsible for innervating the muscles of the tongue. There is only limited information on afferents to the hypoglossal nucleus. A projection has been described from the subcoeruleus nucleus to the hypoglossal nucleus, and this projection may contain some noradrenergic fibers (Aldes, 1990). Indeed, catecholaminergic nerve termi-nals have been shown to synapse with hypoglossal motoneu-rones (Aldes et al., 1990). All three sub-classes of adrenoceptor have been detected in the hypoglossus nucleus: the α1- adreno-ceptors are mainly of the α1A subtype (Day et al., 1997; Pieri-obone et al., 1994), although α1B- and α1D-adrenoceptors are also present (Day et al., 1997), the α2-adrenoceptors are of the α2A subtype (Scheinin et al., 1994), and the β-adrenoceptors are of the β1 subtype (Paschalis et al., 2009). The α1- and β-adrenoceptors are present in high densities, whereas there is only a sparse presence of α2-adrenoceptors. Noradrenaline has been shown to excite hypoglossal neurones via the stimulation of α1-adrenoceptors (Parkis et al., 1995).

4.4.4 Sensory nuclei

4.4.4.1 Sensory trigeminal nuclei. These nuclei, belong-ing to the system of cranial nerve V, receive a rich noradrener-gic input from the LC (Levitt and Moore, 1979; Senba et al., 1981): this pathway has been implicated in antinociception (Cahusac et al., 1995; Couto et al., 2006; Tsuruoka et al., 2003). α1-Adrenoceptors are present in the sensory trigeminal nuclei: they belong to the α1A subtype (Day et al., 1997; Pieribone et al., 1994). α2-Adrenoceptors have been identified in both the mes-encephalic (Scheinin et al., 1994) and spinal (Scheinin et al., 1994; Zhang et al., 1998b) sensory trigeminal nuclei: these α2-adrenoceptors belong to the α2A subtype (Scheinin et al., 1994). The α2-adrenoceptors mediate the inhibitory effect of noradrena-line on spinal trigeminal neurones (Zhang et al., 1998b). Electrical stimulation of the LC has been shown to inhibit pain-sensitive trigeminal neurones (Matsutani et al., 2000; Sasa et al., 1974; Tsu-ruoka et al., 2003).

4.4.4.2 Cochlear nuclei. The cochlear nuclei, together with the vestibular nuclei, belong to the system of cranial nerve VIII. The LC projects to the cochlear nuclei (Klepper and Her-bert, 1991; Kromer and Moore, 1980; Schuerger and Balaban, 1993). α1D-Adrenoceptors have been identified in the cochlear nuclei (Day et al., 1997) and it has been shown that noradrenaline excites cochlear neurones (Ebert, 1996), probably by interacting with these receptors. The noradrenergic projection to the cochlear nuclei suggests that the LC may modulate auditory processing.

4.4.5 Sleep/arousal regulating nuclei. Apart from the LC itself, there are two other major sleep/arousal regulating nuclear complexes in the brainstem: the dorsal raphe (DR) nucleus and the pedunculopontine tegmental (PPT) nucleus/ laterodorsal tegmen-tal (LDT) nucleus. Both these nuclear complexes are under norad-renergic control from the LC (for details, see Section 6.1).

4.4.5.1 DR. The serotonergic neurones of the DR have a wakefulness-promoting function (McGinty and Harper, 1976;

Monti, 2011; Trulson and Jacobs, 1979). The LC sends a noradr-energic projection to the DR (Kim et al., 2004b; Levitt and Moore, 1979; Sakai et al., 1977). All subtypes of α1-adrenoceptor (α1A, α1B and α1D) are present on DR neurones (Day et al., 1997). The LC projection exerts an excitatory influence on DR neurones (Brown et al., 2002), and it has been shown that noradrenaline excites DR neurones via an action at α1-adrenoceptors (Pan et al., 1994; Pudovkina et al., 2003; Yoshimura et al., 1985).

4.4.5.2 PPT and LDT. These nuclei contain cholinergic neurones that are involved both in the promotion of wakefulness and initiation of rapid eye movement (REM) sleep (El Mansaari et al., 1989; Jones, 2005; Kayama et al., 1992; Sakai, 1988). It is likely that there are separate wakefulness-promoting and REM-promoting neurones in the PPT/LDT, the wakefulness-promoting neurones being active during wakefulness and quiescent during REM sleep, and the REM-promoting neurones being quiescent during wakefulness and active during REM sleep (Pal and Mal-lick, 2007). There is a noradrenergic projection from the LC to the pontine nuclei (Nieuwenhuys, 1985) that may exert differential effects on the two populations of cholinergic neurones: the wake-fulness-promoting neurones are excited via the stimulation of α1-adrenoceptors, while the REM-promoting neurones are inhibited via the stimulation of α2-adrenoceptors (Bay et al., 2006; Hou et al., 2002) (for details, see Section 6.1)

4.5 Cerebellum

The cerebellum is responsible for motor coordination that involves the timing, force and extent of muscle contractions (Kiernan, 2005). The LC projects profusely to wide areas of the cerebellum, and in particular to the cerebellar cortex (coeruleo-cerebellar pathway: Olson and Fuxe, 1971; Saigal et al., 1980). All sub-classes of adrenoceptor have been detected in the cerebellar cor-tex. The cerebellar α1-adrenoceptors (Schambra et al., 2005; Stone et al., 2006) belong mainly to the α1A subtype (Day et al., 1997), the α2-adrenoceptors to the α2A subtype (Scheinin et al., 1994), and the β-adrenoceptors to the β1 subtype (Paschalis et al., 2009). The importance of the noradrenergic input for cerebellar function is highlighted by the observation that depletion of the cerebellum of noradrenaline leads to motor impairment (Watson and McElligott, 1984).

4.6 Spinal cord

The LC sends a descending projection (coeruleo-spinal pathway) to all three neuronal pools of the spinal cord: to sensory neurones in the dorsal horn; motoneurones in the ventral horn; and pregan-glionic sympathetic neurones in the IML (Hancock and Fougerousse, 1976; Leong et al., 1984; Proudfit and Clark, 1991; Sluka and Westlund, 1992).

4.6.1 Dorsal horn. The sensory neurones of the dorsal horn are involved in the detection of pain, temperature, touch, and posi-tion/movement (proprioception) (Kiernan, 2005). The LC sends a profuse projection to this compartment of the spinal cord (Frit-schy and Grzanna, 1990; Proudfit and Clark, 1991; Westlund and Coulter, 1980), indicating that the LC can modulate the process-ing of sensory information. The predominant adrenoceptor in this

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

670 Journal of Psychopharmacology 27(8)

region is of the α2 subtype (Shi et al., 1999; Smith et al., 1995). Presynaptic inhibitory α2-adrenoceptors have been identified on excitatory peptidergic interneurones in the dorsal horn, and have been implicated in the inhibition of spinal sensory processing by the LC (Olave and Maxwell, 2002). The modulation of pain sen-sation (nociception) by the LC is well documented (for reviews, see Ossipov et al., 2010; Pertovaara and Almeida, 2006; Willis and Westlund, 1997). It has been shown that electrical stimulation of the LC leads to the inhibition of nociception (Margalit and Segal, 1979; West et al., 1993). Furthermore, α2-adrenoceptor agonists (e.g. clonidine) produce analgesia (Eisenach et al., 1996; Sawynok and Reid, 1986), whereas α2-adrenoceptor antagonists cause hyperalgesia (Green et al., 1998).

4.6.2 Ventral horn. Apart from motoneurones in the brain-stem (see Section 4.4.3), innervating the muscles of the face and head, the LC also projects to motoneurones in the ventral horn of the spinal cord (Jones and Yang, 1985; Nygren and Olson, 1977; Westlund et al., 1983), innervating the skeletal muscula-ture. Excitatory α1-adrenoceptors are present on spinal moto-neurones (Day et al., 1997; Pieribone et al., 1994; Smith et al., 1999) and, via the stimulation of these receptors, the LC exerts a facilitatory action on motoneurone activity (White et al., 1991). In fact, LC activity seems to be essential for the mainte-nance of muscle tone: the suspension of LC activity which occurs during REM sleep (Gottesmann, 2011), and attacks of cataplexy (Wu et al., 1999), leads to total atonia (Peever, 2011).

4.6.3 Intermediolateral cell column. Sympathetic pregangli-onic (cholinergic) neurones are located in the intermediolateral cell column (IML) of the thoracolumbar segments (T1–L2) of the spinal cord, and project to sympathetic ganglia innervating spe-cific target organs (Hamill, 1996). These neurones form distinct clusters (nuclei): this organization enables the selective activa-tion of distinct areas of the sympathetic nervous system, as opposed to generalized sympathetic activation (Appel and Elde, 1988). Although the LC projects to the sympathetic pregangli-onic neurones (Pacák and Palkovits, 2001), the main innervation of these neurones derives from the A5/A7 pontine noradrenergic nuclei (Jones and Yang, 1985; Loewy et al., 1979; Westlund et al., 1983, 1984). Noradrenaline exerts mainly an excitatory effect on these neurones (Lewis and Coote, 1990; Yoshimura et al., 1987), via the activation of α1-adrenoceptors (Day et al., 1997; Pieribone at al., 1994).

Parasympathetic preganglionic (cholinergic) neurones are located in the IML of the sacral segments (S2–S4) of the spinal cord, and project to parasympathetic ganglia embedded in the walls of pelvic target organs or located in their proximity (Hamill, 1996). There is a dense noradrenergic projection from the LC to the sacral IML (Jones and Yang, 1985; Sasa and Yoshimura, 1994; Westlund and Coulter, 1980; Westlund et al., 1984; Yoshimura et al., 1990a, 1990b). Both α1-adrenoceptors (Yoshimura et al., 1990a) and α2-adrenoceptors (Smith et al., 1995; Unnerstall et al., 1984; Yaïci et al., 2002) have been described in the IML of the sacral spinal cord. As there is an intricate intrinsic neuronal net-work in this area, involving both excitatory and inhibitory interneurones (Ranson et al., 2006; Shefchyk, 2001), the cellular localization of the adrenoceptors is likely to be complex. Thus parasympathetic inhibition could result from the direct stimula-tion of inhibitory α2-adrenoceptors on the preganglionic neurones themselves, or may arise indirectly via the stimulation of excitatory

α1-adrenoceptors on inhibitory GABAergic interneurones that synapse with the preganglionic neurones. Alternatively, the stimu-lation of α1-adrenoceptors on excitatory interneurones synapsing with the preganglionic neurones would lead to the excitation of the preganglionic neurones themselves (Yoshimura et al., 1990a). The noradrenergic output from the LC to the IML of the sacral spinal cord has been implicated in the control of pelvic reflexes, such as micturition (Sasa and Yoshimura, 1994; Yoshimura et al., 1990b) and penile erection (Giuliano and Rampin, 2000; Yaïci et al; 2002).

5. Afferents of the LCFigure 5 shows a schematic overview of LC afferents.

The general pattern of inputs to the LC is that of reciprocal innervation: most structures to which the LC projects also send outputs to the LC (for comprehensive reviews of afferents of the LC, see Aston-Jones et al., 1991b; Cederbaum and Aghajnaian, 1978; Luppi et al., 1995: for a review of the neurochemistry of LC afferents, see Singewald and Philippu, 1998).

5.1 Neocortex

Although most areas of the neocortex project to the LC (Cederbaum and Aghajanian, 1978; Luppi et al., 1995), the pre-frontal cortex provides the strongest reciprocal link with it (Arnsten and Goldman-Rakic, 1984; Sesack et al., 1989). There is a powerful excitatory connection from the prefrontal cortex to the LC mediated by excitatory amino acids (Jodo and Aston-Jones, 1997). It has been proposed that this excitatory input may provide tonic activation of the LC (Jodo et al., 1998).

5.2 Amygdala

There is a strong reciprocal connection between the LC and the amygdala: the LC sends a dense projection to the central and basal nuclei of the amygdala (see Section 4.2.1), and receives an input from the central nucleus of the amygdala (Cederbaum and Aghajanian, 1978; Charney et al., 1998; Wallace et al. 1989). Corticotrophin-releasing factor (CRF) has been impli-cated as the neurotransmitter in the pathway from the amyg-dala to the LC (Van Bockstaele et al., 1998b). As the amygdala is the key structure in processing fear and anxiety (Canteras et al., 2010; Charney et al., 1998; Pare and Duvarci, 2012), the two-way connection between the LC and the amygdala under-lies the involvement of the LC in fear and anxiety. On the one hand, the activation of the LC, via the stimulation of the amyg-dala, leads to anxiety (McDougle et al., 1995; Redmond and Huang, 1979). On the other hand, stressful and fear-inducing stimuli (including conditioned fear) via the activation of the amygdala, lead to increased LC activity (Charney and Redmond, 1983; Cullinan et al., 1995; Rasmussen and Jacobs, 1986; Redmond and Huang, 1979).

5.3 Hypothalamus

All three arousal-modulating nuclei of the hypothalamus (VLPO, LH/PF, tuberomammillary nucleus (TMN)) project to the LC, consistent with the central role of the LC in the sleep-arousal net-work (see Section 6.1).

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 671

5.3.1 VLPO. The VLPO is a major sleep-promoting nucleus, containing GABAergic inhibitory neurones (Szymusiak and McGinty, 2008). The VLPO sends an inhibitory projection to the LC (Cederbaum and Aghajanian, 1978; Lee et al., 2005a; Peyron et al., 1995; Steininger et al., 2001). This projection forms one limb of a reciprocal inhibitory connection between the VLPO and the LC: the LC inhibits the VLPO via an inhibitory norad-renergic output involving the stimulation of α2-adrenoceptors on VLPO neurones (see Section 4.3.2.1). During slow wave sleep, the GABAergic VLPO neurones are active (Szymusiak and McGinty, 2008), and inhibit the activity of noradrenergic

neurones in the LC (Gervasoni et al., 1998; Nitz and Siegel, 1997). On the other hand, during wakefulness, the LC neurones are active, and inhibit the sleep-promoting neurones of the VLPO (for details, see Section 6.1).

5.3.2 LH/PF. The LH/PF is a major wakefulness-promoting nucleus, containing orexinergic (hypocretinergic) excitatory neu-rones (De Lecea et al., 1998; Szymusiak and McGinty, 2008). The LH/PF densely innervates the neurones of the LC (Date et al., 1999; España et al., 2005; Horvath et al., 1999; Lee et al., 2005a). This pathway has been shown to activate the LC, leading to an

Figure 5. Afferents of the locus coeruleus (LC). Arrows indicate projections. Red arrow: excitatory; blue arrow: inhibitory. Lettering next to arrow indicates neurotransmitter involved: Coloring of source area indicates function (for code see figure). 5-HT: 5-hydroxytryptamine (serotonin); A: adrenaline; A (in circle): amygdala; ACh: acetylcholine; CRF: corticotrophin releasing factor; DA: dopamine; DR: dorsal raphe; Enc: encephalin; GABA: γ-aminobutyric acid; Glu: glutamate; H: histamine; LH: lateral hypothalamic area; Ox: orexin; PAG: periaqueductal grey; PPT/LDT: pedunculopontine tegmental/laterodorsal tegmental nuclei; PrH: nucleus prepositus hypoglossi (dorsomedial rostral medulla); PVN: paraventricular nucleus; RVLM: rostroventrolateral medulla; TMN: tuberomammillary nucleus; V: sensory trigeminal nuclei; VLPO: ventrolateral preoptic area; VP: vasopressin; VTA: ventral tegmental area. Sleep-promoting neurones in the hypothalamus promote slow wave sleep (SWS), and in the brainstem rapid eye movement (REM) sleep. The four nuclear groups in the RVLM are REM-sleep-promoting GABAergic neurones; premotor sympathetic/wakefulness-promoting glutamatergic neurones, premotor sympathetic adrenergic neurones, and pain-modulating encephalinergic sensory relay neurones. The amygdala sends an excitatory projection to the LC. Projections of the wakefulness-promoting neurones in the hypothalamus (LH, TMN) and the brainstem (VTA, PAG, PT/LDT, DR) have excitatory effects, whereas projections of sleep promoting neurones in the hypothalamus (VLPO) and REM-sleep-active neurones in the brainstem (PrH and GABArgic neurones of RVLM) have inhibitory effects on the LC. Premotor sympathetic neurones in the hypothalamus (PVN) and the brainstem (glutamatergic neurones in the RVLM) stimulate the LC, whereas adrenergic premotor sympathetic neurones in the RVLM inhibit the LC. The glutamatergic projection from the RVLM to the LC, by activating the LC, also has a wakefulness promoting function (see Section 5.4.2.1). Trigeminal and spinal sensory neurones evoke excitatory effects, whereas encephalinergic sensory relay neurones in the RVLM have inhibitory effects on the LC.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

672 Journal of Psychopharmacology 27(8)

increase in arousal and suppression of REM sleep (Bourgin et al., 2000; Hagan et al., 1999). It has been shown recently that the orexinergic input from the LH/PF activates a subgroup of LC neu-rones that project to the medial prefrontal cortex, and it has been suggested that this pathway may malfunction in narcolepsy, lead-ing to excessive daytime sleepiness (Del Cid-Pellitero and Garzn, 2011).

5.3.3 TMN. The TMN is a wakefulness-promoting nucleus, containing histaminergic neurones (Szymusiak and McGinty, 2008). An excitatory projection from the TMN to the cerebral cor-tex, involving the stimulation of H1 histamine receptors, has been implicated in the wakefulness-promoting function of this nucleus (Haas and Panula, 2003). There is a projection from the TMN to the LC (Iwase et al., 1993; Lee et al., 2005b), which exerts an excitatory effect via the activation of H1 histamine receptors (Korotkova et al., 2005). In this way, the direct wakefulness-pro-moting effect of the TMN via its cortical projection is likely to be augmented indirectly via the activation of the LC. There is no strong direct reciprocal connection between the TMN and the LC, since the noradrenergic innervation of the TMN derives mainly from nuclei A1/A2, and the projection from the LC to this nucleus is rather weak (Ericson et al., 1989). However, the LC exerts a powerful indirect influence on TMN activity via inhibition of the VLPO (see Section 6.1).

5.3.4 PVN. The PVN is an important premotor sympathetic nucleus which projects directly to preganglionic sympathetic neu-rones in the spinal cord (Hosoya et al., 1991; Ranson et al., 1998) via vasopressinergic fibers (Pyner, 2009). The PVN also projects indirectly to spinal preganglionic sympathetic neurones via the LC. The LC receives a dense projection from the PVN (Luiten et al., 1985; Luppi et al., 1995; Reyes et al., 2005). This projection consists of CRF-containing nerve fibers, which exert an excitatory influence on noradrenergic neurones in the LC (Reyes et al., 2005). Thus the PVN potentiates the sympathoexcitatory effect of the LC. On the other hand, the LC facilitates PVN activity via a reciprocal excitatory projection to this nucleus (see Section 4.3.2.3).

5.4 Brainstem

5.4.1 Midbrain and pons. Apart from the LC itself, there are four nuclear groups in the upper brainstem that are intimately linked to sleep/arousal mechanisms. All these four nuclear groups project to the LC.

5.4.1.1 Ventral tegmental area. Apart from their well-established roles in reward, motivation and drug addiction (Nicola et al., 2005; Wise, 2004), the dopaminergic neurones of the ven-tral tegmental area (VTA) are also associated with a wakefulness-promoting function (Monti and Jantos, 2008; Monti and Monti, 2007; Niepel et al., 2013). The VTA projects to the LC (‘meso-coerulear pathway‘) (Beckstead et al., 1979; Ornstein et al., 1987; Simon et al., 1979a, 1979b), and there is evidence that this path-way exerts an excitatory influence on LC neurones (Deutch et al., 1986). Thus, the wakefulness-promoting function of the VTA may be mediated, at least partly, by the LC. Indeed, it has been shown that dopamine applied to the LC inhibits sleep (Crochet and Sakai, 2003).

5.4.1.2 Periaqueductal grey matter. The periaqueductal grey matter (PAG) consists of a heterogeneous group of neurones surrounding the cerebral aqueduct in the midbrain. The PAG has been implicated in a number of diverse functions, such as pain control (Heinricher et al., 2009), fear and anxiety (Canteras et al., 2010), arousal (Monti and Jantos, 2008), REM sleep (Luppi et al., 2012) and cardiovascular control (Dampney et al., 2013). Inter-estingly, the central noradrenergic system is also involved in the regulation of these same functions (see Sections 4.1.1, 4.6.1 and 6.1). The ventrolateral PAG receives a noradrenergic input from the A1 and A2 groups of neurones (Herbert and Saper, 1992), and in turn projects to the A5 (Bajic et al., 2012) and A7 (Bajic et al., 2001) noradrenergic nuclei. There is evidence that the LC (A6 group of noradrenergic neurones) also receives an input from neurones in the PAG (Bajic and Proudfit, 1999; Cameron et al., 1995; Ennis et al., 1991; Lee et al., 2005a; Luppi et al., 1995; Van Bockstaele et al., 2001). Neurones in the ventrolateral PAG project to the ventromedial aspect of the peri-LC (Van Bockstaele et al., 2001) where they synapse with the dendrites of LC neurones (Bajic et al., 2000). The projection from the ventrolateral PAG to spinally projecting noradrenergic neurones is involved in pain modulation: stimulation of neurones in the PAG produces antin-ociception (Bajic and Proudfit, 1999). A degree of topographic functional specifity of the projections from the PAG to noradren-ergic nuclei has been suggested: the projection to the A7 nucleus may play a role mainly in analgesia (Bajic and Proudfit, 1999; Bajic et al., 2001) whereas the projection to the A5 nucleus and the LC has been implicated in cardiovascular regulation (Bajic and Proudfit, 1999; Bajic et al., 2012). A group of wakefulness-active dopaminergic neurones have been identified in the ventro-lateral PAG (Lu et al., 2006): these neurones may contribute to the activation of the LC during wakefulness which, in turn, could lead to the augmentation of dopaminergic arousal (Niepel et al., 2013).

5.4.1.3 DR. The DR is a wakefulness-promoting nucleus, and the serotonergic neurones of the DR are active during wake-fulness (Monti, 2011; Monti and Jantos, 2008). DR neurones pro-ject to the LC (Kim et al., 2004b; Luppi et al., 1995; Pasquier et al., 1977; Vertes and Kocsis, 1994), forming one limb of the reciprocal connection between the two nuclei.

5.4.1.4 PPT/LDT. The cholinergic neurones of the PPT/LDT perform a dual role in the sleep/arousal network: some of these neurones are wakefulness promoting (Datta et al., 2011), whereas others are involved in the initiation/mediation of REM sleep (Jones, 1991; Xi et al., 2012). The PPT/LDT neurones project to the LC (Jones and Yang, 1985). Cholinergic nerve terminals are detectable in the LC (Jones, 1990), and acetylcho-line has an excitatory effect on noradrenergic neurones in the LC (Egan and North, 1985; Engberg and Svenson, 1980). The cholinergic output from the PPT/LDT also seems to facilitate noradrenergic neurotransmission in target areas, such as the VLPO (Saint-Mleux et al., 2004) and the DR (Li et al., 1998), by stimulating release-modulating nicotinic receptors on noradr-energic nerve terminals.

5.4.2 Medulla oblongata. The medulla contains a large num-ber of varied neuronal groups involved in the regulation of a num-ber of physiological functions, such as autonomic activity, cardiovascular regulation, respiration, thermoregulation and pain processing (Goodchild and Moon, 2009; Heinricher et al., 2009; Kumagai et al., 2012; Morrison and Nakamura, 2011; Van

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 673

Bockstaele and Aston-Jones, 1995), and many of these neuronal groups project to the LC. It is beyond the scope of this review to cover all these connections in detail, however, some of the most important ones are described below.

5.4.2.1 RVLM. The RVLM, an important premotor sympa-thetic area (see Sections 4.4.1.1 and 6.2), contains a number of distinct subnuclei with distinct neurochemical identities and pro-jection patterns. One group of neurones is identified as the nucleus paragigantocellularis (PGi), another one as the C1 nucleus of adr-energic neurones (Astier et al., 1990). Neurones in the RVLM project to the LC (Astier et al., 1990; Aston-Jones et al., 1986; Guyenet and Young, 1987; Van Bockstaele and Aston-Jones, 1992; Van Bockstaele et al., 1998a). Interestingly, some neurones in the RVLM send collaterals both to the LC and spinal pregan-glionic sympathetic neurones (Van Bockstaele and Aston-Jones, 1995).

There is a major glutamatergic output from the RVLM to the LC which exerts an excitatory influence on LC neurones (Ennis and Aston-Jones, 1986, 1988; Van Bockstaele and Aston-Jones, 1995). This projection activates the LC, and the glutamatergic neurones have the function of premotor sympathetic neurones. It has been suggested that the activation of the LC by the RVLM may contribute to the wakefulness-promoting function of the LC (Van Bockstaele and Aston-Jones, 1995). Indeed, it has been reported that stimulation of inhibitory α2-adrenoceptors in the RVLM, evoked by the local injection of clonidine, leads to seda-tion (Yamazato et al., 2001).

Adrenergic neurones in the C1 nucleus within the RVLM pro-ject to the LC, and exert an inhibitory effect on LC neurones via the stimulation of α2-adrenoceptors (Aston-Jones et al., 1992; Van Bockstaele and Aston-Jones, 1995).

Inhibitory GABAergic neurones in the RVLM send efferent fibers to the LC: some of these neurones are active during REM sleep, and may play a role in switching off the LC during REM sleep (Sirieix, 2012).

The RVLM also contains enkephalinergic neurones (Johnson et al., 2002) which project to the LC (Drolet et al., 1992). This projection activates opiate receptors on LC neurones (Van Bockstaele, 1998) which leads to the inhibition of neuronal firing (Illes and Norenberg, 1990; Pepper and Henderson, 1980). The LC has been implicated in opiate-induced analgesia (Drolet et al., 1992) and opiate withdrawal (Johnson et al., 2002).

5.4.2.2 Dorsomedial rostral medulla. In this area of the medulla the nucleus prepositus hypoglossi (PrH) has been identi-fied as a distinct group of neurones projecting to the LC (Aston-Jones et al., 1986). The PrH neurones are GABAergic, and exert an inhibitory action on the LC (Ennis and Aston-Jones, 1989a, 1989b). The GABAergic neurones in the PrH have been impli-cated in the regulation of REM sleep: the inhibitory output to the wakefulness-promoting LC neurones may be responsible for the silencing of these neurones during REM sleep (Kaur et al., 2001). The involvement of the PrH in REM sleep is supported by its reciprocal connection to the PPT (Higo et al., 1990) which has a role in the generation of REM sleep that is well established (see Sections 4.4.5.2 and 5.4.1.4). It should be noted, however, that other GABAergic medullary nuclei, sending projections to the LC, such as the dorsal paragigantocellular reticular nucleus (Goutagny et al., 2008) and the lateral paragigantocellular nucleus (see Sec-tion 5.4.2.1), have also been implicated in the neuronal network responsible for the generation and regulation of REM sleep.

5.4.3. Sensory trigeminal nuclei. The three sensory trigemi-nal nuclei stretch through the whole length of the brainstem: the principal sensory nucleus is located in the pons, the mesence-phalic nucleus occupies the midbrain, and the spinal nucleus extends through the whole length of the medulla as far as the dor-sal column of the spinal cord (Kiernan, 2005). The sensory tri-geminal nuclei project to the LC (Couto et al., 2006; Craig, 1992), providing the anatomical basis for the processing and modulation of pain signals from the face and head. It has been shown that noxious stimulation of the trigeminal sensory area (e.g. electrical tooth pulp stimulation or noxious air puff to the face) activates the LC (Kimura and Nakamura, 1985; Singewald et al., 1999; Voisin et al., 2005). Furthermore, the noxious stimulation-evoked LC activation leads to concomitant neuronal activation in the somato-sensory thalamus, consistent with the processing of pain signals via the LC (Voisin et al., 2005).

5.5 Spinal cord

Sensory neurones in the dorsal horn of the spinal cord send an ascending projection to the LC (Cederbaum and Aghajanian, 1978; Craig, 1992). This pathway has been implicated in the detection and processing of thermal and nociceptive stimuli. There is extensive experimental evidence showing that noxious and/or thermal stimulation leads to the activation of the LC (Elam et al., 1986; Hirata and Aston-Jones, 1994; Kimura and Nakamura, 1985; Palkovits et al., 1995; Rasmussen et al., 1986; Sugiyama et al., 2012).

6. The LC in regulatory networksThe full functional significance of the connections of the central noradrenergic neurones, and in particular of the LC, becomes apparent when the integrated neuroanatomy of physiological reg-ulatory neural networks is considered. Such networks include the sleep/arousal network, the autonomic network and the network related to the modulation of physiological functions by light (‘photomodulation‘). The LC also plays a role in some other important networks, such as the fear/anxiety and the pain modula-tion networks (see Section 8; Samuels and Szabadi, 2008b; Szabadi, 2012).

6.1 Sleep/arousal network

This neural network is responsible for the regulation of arousal. Three states of arousal (also referred to as vigilance states, e.g. Fort et al., 2009) can be distinguished: wakefulness, characterized by fast, small-amplitude, desynchronized EEG, active muscle tone and conscious awareness; slow wave sleep (non-REM sleep), characterized by syn-chronized, slow, large-amplitude EEG, reduced muscle tone, lack of conscious awareness; REM sleep (paradoxical or active sleep), charac-terized by fast, desynchronized EEG, loss of muscle tone, and variable levels of conscious awareness (Jones, 2005; Lin et al., 2011). The sleep/arousal network consists of distinct wakefulness-promoting and sleep-promoting nuclei, and their connections, located in the diencephalon (thalamus and hypothalamus) and the brainstem: the wakefulness-pro-moting nuclei are active during wakefulness and quiescent during sleep, whereas the sleep-promoting nuclei are active during sleep and quiescent during wakefulness (Aston-Jones, 2005; De Andrés et al.,

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

674 Journal of Psychopharmacology 27(8)

2011; Fuller et al., 2006; Haas and Lin, 2012; Jones, 2005; Lin et al., 2011; Schwartz and Roth, 2008; Szabadi, 2006; Szymusiak and McGinty, 2008). The three vigilance states, and the transitions between them, are regulated by the subcortical sleep/arousal network. During slow wave sleep the cerebral cortex displays its synchronized intrinsic activity: wakefulness and/or REM sleep ensue when this ‘cortical solil-oquy‘ is disrupted by afferent inputs from subcortical areas, such as the diencephalon and the brainstem (Lin et al., 2011). The main role of the GABAergic sleep-promoting neurones is to switch off the wakeful-ness-promoting neuronal groups and allow sleep to develop (Lin et al., 2011) although the GABAergic sleep-promoting neurones of the BF exert a direct inhibitory influence on cortical excitability (see Section 4.1.2). A separate network has been implicated in the regulation of REM sleep, involving complex interactions between GABAergic, glu-tamatergic and cholinergic neurones in the brainstem (Fort et al., 2009; Fuller et al., 2007; Vetrivelan et al., 2011): however, a detailed discus-sion of this network is beyond the scope of this review.

A schematic diagram of the sleep/arousal network is shown in Figure 6. All the wakefulness-promoting nuclei project directly to the cerebral cortex. These include, in the diencephalon, the glutamatergic neurones of the medial and intralaminar nuclei of the thalamus (Jones, 2005; Lin et al., 2011), and, in the hypothalamus, the histaminergic

neurones of the TMN and the orexinergic neurones of the LH/PF (Haas and Lin, 2012). The wakefulness-promoting nuclei in the brain-stem are the serotonergic DR (Monti, 2011), the dopaminergic neu-rones of the periaqueductal grey and the VTA (Monti and Jantos, 2008; Niepel et al., 2012), and some of the cholinergic neurones of the pedunculopontine/laterodorsal nuclei (Jones, 2005). The LC is a major wakefulness-promoting nucleus in the brainstem, sending a powerful excitatory projection to the cerebral cortex (see Section 4.1.1). The major sleep-promoting nucleus is the VLPO of the hypothalamus: GABAergic neurones in this nucleus exert a powerful inhibitory influ-ence on the TMN and the LC, and thus promote sleep.

The LC occupies a central position in the hub of the sleep/arousal network, where it fulfils four wakefulness-promoting functions. Firstly, it evokes a direct alerting effect via its cortical projection. Secondly, it augments the alerting influence of other wakefulness-promoting nuclei by sending excitatory projections to them (see Sections 4.1.2, 4.3.1, 4.4.5.1, 4.4.5.2) (connections to BF and DR are not shown in Figure 6). Thirdly, it receives inputs from other wakefulness-promoting nuclei (see Sections 5.3.2, 5.3.3, 5.4.1.1, 5.4.1.2, 5.4.1.3, 5.4.1.4), which potentiate the alert-ing effect of the LC. Fourthly, by inhibitory projections it switches off the activities of sleep-promoting neurones in the BF (see

Figure 6. Schematic diagram of the connections within the arousal-controlling neuronal network. γ-Aminobutyric acid (GABA)ergic interneurones, in (white). LC: locus coeruleus; LH/PF: lateral hypothalamic/perifornical area; PPT: pedunculopontine tegmental nucleus; R: raphe nuclei; Th: thalamus; TMN: tuberomammillary nucleus; VLPO: ventrolateral preoptic nucleus; VTA: ventral tegmental area. Neuronal outputs: excitatory (red arrows); inhibitory (blue arrows). Neurotransmitters: 5-HT: 5-hydroxytryptamine; 5-HT2A and 5-HT2C: excitatory 5-HT receptors; ACh: acetylcholine; DA: dopamine; Glu: glutamate; H: histamine; H1: excitatory H1 histamine receptors; NA: noradrenaline; Ox: orexin. Receptors: α1: excitatory α1-adrenoceptors; α2: inhibitory α2-adrenoceptors. Wakefulness-promoting nuclei (yellow), sleep-promoting nucleus (purple). The wakefulness-promoting nuclei exert a direct activating effect on the cerebral cortex: the VLPO promotes sleep by inhibiting the TMN and the LC. The LC promotes wakefulness by stimulating the cerebral cortex and the wakefulness-promoting neurones of the PPT, and by inhibiting the VLPO. The LC also inhibits the REM-sleep-promoting neurones of the PPT. The raphe nucleus promotes wakefulness by activating the cerebral cortex; this effect is attenuated by stimulation of GABAergic interneurones, which inhibit the LC and the VTA. The GABAergic interneurones, activated by excitatory 5-HT2C receptors, are located in the VTA itself (Bubar and Cunningham, 2007; Gobert et al., 2000) and in the vicinity of the LC (Gobert et al., 2000). The VTA exerts its wakefulness-promoting effect largely via activation of the LC, and the LH/PF largely via activation of the TMN and the LC. See Section 6.1 for details. Reproduced from Samuels ER and Szabadi E (2008b) Functional neuroanatomy of the noradrenergic locus coeruleus: Its roles in the regulation of arousal and autonomic function. Part II: Physiological and pharmacological manipulations and pathological alterations of locus coeruleus activity in humans. Curr Neuropharmacol 6: 254–285 with permission from Bentham Science Publishers.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 675

Figure 4 and Section 4.1.2) and the VLPO (see Section 4.3.2.1), and of REM-promoting activities of some of the REM-active neu-rones in the PPT/LDT (see Section 4.4.5.2).

Further details of the sleep/arousal network are described in the legend to Figure 6. The interplay between excitatory choliner-gic and inhibitory GABAergic neurones in the BF, and their mod-ulation by the LC, are shown in Figure 4. Recent work has emphasized the key role of the BF, and its activation by a gluta-matergic input from the parabrachial nucleus and the subcoer-uleus area, in the maintenance of arousal (Fuller et al., 2011).

6.2 Autonomic network

The autonomic nervous system supplies smooth muscle, cardiac muscle and glandular tissue with vegetative motor fibers, and influences metabolic processes via the regulation of the release of hormones (Loewy, 1990a). Both divisions (sympathetic and para-sympathetic) of the autonomic nervous system consist of two neu-rones linked serially: a preganglionic neurone located in the CNS (brainstem or spinal cord) synapses with a postganglionic neurone located in a peripheral autonomic ganglion. Both the sympathetic and parasympathetic preganglionic neurones are cholinergic. The sympathetic postganglionic neurones are noradrenergic, with the exception of those innervating sweat glands (‘sudomotor neu-rones‘), which are cholinergic: all parasympathetic postganglionic neurones are cholinergic (Hamill, 1996; Loewy, 1990a). The activity of preganglionic neurones is controlled by a number of premotor autonomic nuclei (e.g. CR, LC (together with the A5/A7 pontine noradrenergic nuclei), PVN, RVLM) (Loewy, 1990b; Mosqueda-Garcia, 1996), which in turn are under the influence of ‘higher‘ structures, such as the suprachiasmatic nucleus (SCN) of the hypothalamus (Buijs et al., 2003), the amygdala (Fortaleza et al., 2012) and the cerebral cortex (Cechetto and Saper, 1990).

The pontine noradrenergic nuclei (A5/A7, LC) occupy a cen-tral position in the autonomic network (Figures 7 and 8), where they accomplish three separate functions. Firstly, they are impor-tant sympathetic premotor nuclei, exerting an excitatory influence on preganglionic sympathetic neurones in the IML via the stimu-lation of α1-adrenoceptors (see Section 4.6.3). Secondly, they also function as premotor parasympathetic nuclei: they project to preganglionic parasympathetic nuclei in the brainstem, such as the EWN, the salivatory and vagal nuclei, where they exert an inhibi-tory influence via the stimulation of α2-adrenoceptors (see Section 4.4.2). Thirdly, the LC modulates the activity of other sympathetic premotor nuclei, such as the CR, PVN and RVLM (see Sections 4.3.2.3 and 4.4.1). While the effect of the LC on PVN and CR neurones is excitatory via the activation of α1-adrenoceptors, the LC exerts an inhibitory influence on the vasomotor and cardio-accelerator neurones of the RVLM via the stimulation of inhibi-tory α2-adrenoceptors, thereby presumably dampening the sympathoexcitation evoked by sympathetic premotor nuclei. The central role of the LC in the control of cardiac activity is shown in Figure 8.

On the basis of the position of the pontine noradrenergic nuclei within the autonomic network, it is possible to predict changes in autonomic activity resulting from their activation. Overall, central noradrenergic activation leads to sympathoexcitation and para-sympathetic inhibition. The signs of sympathoexcitation are an increase in pupil diameter (Phillips et al. 2000a, 2000b) and mod-erate increases in heart rate and blood pressure (Drolet and

Gauthier, 1985; Stock et al., 1981) whereas the signs of parasym-pathetic inhibition are the attenuation of the light reflex response (Bakes et al., 1990; Bitsios et al., 1996) and a reduction in salivary flow (Szabadi and Tavernor, 1999).

As the pontine noradrenergic nuclei are intimately involved in the control of pupil function, it is expected that changes in LC activity would be reflected in changes in pupil diameter. Indeed, it has been shown that fluctuations in the firing rate of LC neurones are closely paralleled by fluctuations in pupil diameter (Aston-Jones and Cohen, 2005; Rajkowski et al., 1993). Furthermore, as the LC integrates arousal-related information from all wakeful-ness-promoting and sleep-promoting nuclei (see Section 6.1 and Figure 6), it is expected that LC activity at any one time would reflect the overall level of arousal, and this could be reflected in pupillary function. The relationship between level of arousal and pupillary function can be assessed by the Pupillographic Sleepiness Test (PST), based on an inverse relationship between level of arousal and pupillary fluctuations in darkness (Hou et al., 2006, 2007; Wilhelm et al., 2001; Yoss et al., 1970). It is an intriguing possibility that, as sleepiness develops, there is an increasing instability between wakefulness-promoting and sleep-promoting inputs to the LC, leading to increasing fluctuations in LC neuronal activity, and this may be detected by the PST.

6.3 Photomodulation network

Photoreceptors in the retina are responsible for two separate functions: pattern recognition and irradiance detection. Pattern recognition leads to image formation (IF), via the process of visual perception, whereas irradiance detection gauges the intensity of ambient light, modulating a number of non-image formation (NIF) functions. NIF functions include the pupillary light reflex, adjustment of the circadian cycle to the day/night cycle (photoentrainment), and acute influence of light on some physiological functions (e.g. arousal, sympathetic activity). The term ‘photomodulation‘ is used here to refer to this latter effect of light.

It is well documented that the stimulation of the photoreceptor cells of the retina, the rods and cones, leads to the activation of retinal ganglion cells, that in turn give rise to both IF and NIF projections. The IF output is to the visual cortex, via the lateral geniculate nucleus, whereas the two major NIF outputs are to the hypothalamus and the diencephalon/midbrain border (OPN) (see below).

It has been discovered recently that the rods and cones are not the only photosensitive cells in the retina: a small population of retinal ganglion cells scattered randomly over the whole area of the retina are also photosensitive (ipRGCs). It has also been shown that the photosensitivity of these ganglion cells is due to the presence of the photo-pigment melanopsin (Semo et al., 2005). Melanopsin has a distinct spectral sensitivity: it is activated by a narrow short-wavelength band (around 460 nm) of blue light (Kawasaki and Kardon, 2007; Thapan et al., 2001).

There is evidence that NIF functions are mediated almost exclusively by the ipRGCs that can be stimulated by light, both directly and indirectly, via the rods and cones (Gooley et al., 2003; Güler et al., 2008). The ipRGCs give rise to two pathways: one to the OPN and another one to the hypo-thalamus. Recent evidence indicates that the two pathways originate from separate populations of ipRGCs (Baver et al., 2008; Chen et al., 2011). The OPN is a promotor nucleus in the parasympathetic light reflex pathway: it

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

676 Journal of Psychopharmacology 27(8)

projects to the preganglionic parasympathetic neurones in the EWN (see Section 4.4.2.1). The output to the hypothalamus (retino-hypothalamic tract) innervates two hypothalamic nuclei, the SCN and the VLPO (Gooley et al., 2003; Güler et al., 2008; Lu et al., 1999) (Figure 9). The SCN is responsible for the generation of circadian rhythms and is involved in the modulation of autonomic and endocrine functions (Kalsbeek et al., 2000, 2006; Van Esseveldt et al., 2000) whereas the VLPO is a major sleep-pro-moting nucleus (see Section 6.1 and Figure 6).

Light has a dual effect on arousal and sympathetic activity: it has both wakefulness-promoting and sleep-promoting functions, and sym-pathoexcitatory and sympathoinhibitory functions. The overall effect of exposure to light depends on the relationship between the two

opposing functions. Interestingly, there is a species difference relating to the overall effect of light on arousal: it is sleep-promoting in noctur-nal animals (Altimus et al., 2008; Tsai et al., 2009) whereas in diurnal animals, including man, it is wakefulness-promoting (Cajochen et al., 2005; Lockley et al., 2006; Revell et al. 2006). The overall effect of light, in both diurnal and nocturnal animals, is inhibitory on pregangli-onic sympathetic neurones of the upper thoracic spinal cord projecting to the superior cervical ganglion (Kalsbeek et al., 2000; Pereau-Lenz et al., 2003): this effect leads to pupil constriction (Passatore, 1976; Passatore and Pettorossi, 1976; Szabadi et al., 2010) and inhibition of melatonin synthesis (Kalsbeek et al., 1999; Nishino et al., 1976; Zeitzer et al., 2000). However, the overall effect of light, in diurnal animals, is

Figure 7. Noradrenergic regulation of autonomic functions. Outputs from the nuclei are indicated by arrows. Red: excitatory (+); blue: inhibitory (–).Neurotransmitters: ACh: acetylcholine; NA: noradrenaline. Nuclei: EWN: Edinger-Westphal nucleus; LC: locus coeruleus; Para ggl: parasympathetic ganglion; Preggl Para: preganglionic parasympathetic neurones; Preggl Symp: preganglionic sympathetic neurones; PVN: paraventricular nucleus; Saliv: salivatory nucleus; Symp ggl: sympathetic ganglion. Receptors: α1 and α2: adrenoceptor subtypes. Organs comprising of smooth muscle (e.g. blood vessels, iris) or glandular tissue (e.g. sweat glands, salivary glands) receive autonomic (sympathetic and parasympathetic) innervations. Both innervations consist of a chain of two neurones (preganglionic and postganglionic) joined in a synapse located in the autonomic ganglion. Preganglionic sympathetic neurones are located in the intermediolateral cell column (IML) of the spinal cord, whereas the preganglionic parasympathetic neurones are located in brainstem nuclei. Blood vessels (arterioles) and sweat glands receive sympathetic and salivary glands parasympathetic inputs, whereas the smooth muscles in the iris are controlled by opposing sympathetic and parasympathetic inputs. The preganglionic neurones are always cholinergic: the postganglionic sympathetic neurones are noradrenergic (with the exception of those innervating the sweat glands which are cholinergic) whereas the postganglionic parasympathetic neurones are always cholinergic. The preganglionic neurones are influenced by premotor autonomic nuclei of which three are shown: PVN; rostroventrolateral medulla (RVLM) containing vasomotor and cardioaccelerator neurones; and the LC. The pontine noradrenergic nuclei (LC, A5/A7) play a pivotal role in autonomic regulation, influencing the activities of preganglionic neurones both directly and indirectly via other premotor nuclei. The outputs from the pontine noradrenergic nuclei can activate either excitatory α1-adrenoceptors or inhibitory α2-adrenoceptors. The noradrenergic nuclei exert an excitatory effect on preganglionic sympathetic neurones, and an inhibitory effect on premotor sympathetic neurones (vasomotor neurones) in the RVLM and on preganglionic parasympathetic neurones. The activity of the premotor autonomic neurones (e.g. PVN) is under the influence of the cerebral cortex. The light reflex is a parasympathetically-mediated reflex consisting of the constriction of the pupil in response to a light stimulus reaching the retina. The neuronal chain in the reflex includes the pretectal nucleus, the EWN, and the ciliary ganglion (CG). See Section 6.2, for details. Reproduced from Samuels ER and Szabadi E (2008b) Functional neuroanatomy of the noradrenergic locus coeruleus: Its roles in the regulation of arousal and autonomic function. Part II: Physiological and pharmacological manipulations and pathological alterations of locus coeruleus activity in humans. Curr Neuropharmacol 6: 254–285 with permission from Bentham Science Publishers.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 677

stimulatory at the level of lower thoracic segments, leading to stimula-tion of cardiovascular activity (Cajochen et al., 2005; Michimori et al., 1997; Scheer et al., 1999). On the other hand, light seems to inhibit cardiovascular activity in nocturnal animals (Scheer et al., 2001).

The dual effect of light on arousal and sympathetic activity is controlled by the photomodulation network (Figure 9). The LC occupies an important position in this network: it has a central role in mediating the stimulatory effect of light on arousal and sympathetic activity. The glutamatergic ipRGCs project to the VLPO (Lu et al. 1999): this pathway is respon-sible for mediating the sleep-promoting effect of light. The ipRGCs also project to the SCN where they can stimulate separate populations of GABAergic inhibitory and gluta-matergic excitatory output neurones (Mistlberger, 2005). An inhibitory GABAergic output from the SCN to the PVN, an important premotor sympathetic nucleus, can switch off the sympathoexcitatory drive from the PVN to the preganglionic sympathetic neurones in the IML, leading to

sympathoinhibition (Kalsbeek et al., 2000). On the other hand, an excitatory glutamatergic output form the SCN (Perreau-Lenz et al., 2003) can activate orexinergic neurones in the dor-somedial hypothalamus (DMH) which, in turn, exert an excitatory influence on the LC (Aston-Jones, 2005; Aston-Jones et al., 2001; Gonzalez and Aston-Jones, 2006). Indeed, it has been shown by functional magnetic resonance imaging (fMRI) in humans that short wavelength (blue) light, that pref-erentially stimulates the melanopsin-containing photorecep-tors of ipRGCs (Kawasaki and Kardon, 2007), activates the LC (Vandewalle et al., 2007). The activation of the LC by light via the SCN→DMH→LC route would lead to wakefulness-promoting (see Section 6.1) and sympathetic activating (see Section 6.2) effects. Interestingly, another pathway mediating the stimulatory effect of light on arousal has been described recently: light can activate the orexinergic neurones of the LH/PF area of the hypothalamus which in turn activate wake-fulness-promoting serotonergic neurones in the DR

Figure 8. Schematic diagram of the neuronal network regulating cardiac activity.Cellular groups (nuclei and ganglia): CG: cardiac ganglia; CR: caudal raphe; IML: intermedial lateral column; LC: locus coeruleus; PVN: paraventricular nucleus; RVLM: rostral ventrolateral medulla; StG: stellate ganglion; V: vagal nuclei (dorsal nucleus, nucleus ambiguous). Outputs from nuclei and ganglia: arrows (red, excitatory; blue, inhibitory). Neurotransmitters: 5-HT: 5-hydroxytryptamine (serotonin); ACh: acetylcholine; Glu: glutamate; NA: noradrenaline; VP: vasopressin. Adrenoceptors: α1 (excitatory), α2 (inhibitory), β1 (excitatory). Cardiac activity is under dual sympathetic/parasympathetic control: sympathetic output to the heart stimulates it, whereas the parasympathetic output inhibits it. Sympathetic output: sympathetic preganglionic cholinergic neurones in the (IML) of the spinal cord synapse with postgangli-onic noradrenergic neurones in the stellate ganglion; the postganglionic neurones innervate the heart where they mediate a stimulatory effect (cardioacceleration) via excitatory β1-adrenoceptors. Parasympathetic output: parasympathetic preganglionic cholinergic neurones in the vagal nuclei synapse with postganglionic cholinergic neurones in the cardiac ganglia; the postganglionic neurones innervate the heart where they mediate an inhibitory effect (cardiodeceleration) via inhibitory muscarinic cholinoceptors. Premotor autonomic neurones regulate the activity of preganglionic neurones. The CR, LC, PVN (together with the other pontine noradrenergic nuclei) and RVLM send excitatory outputs to preganglionic sympathetic neurones in the IML, and the LC sends an inhibitory output to preganglionic parasympathetic neurones in the vagal nuclei, thereby promoting sympathoexcitation. The pontine noradrenergic nuclei play a central role in cardiac regulation: apart from controlling the activity of preganglionic sympathetic and parasympathetic neurones, they also modulate the activity of other premotor autonomic nuclei. The LC promotes sympathetic activity via excitatory outputs to the CR and PVN but also attenuates it via an inhibitory output to the RVLM. Further attenuation of sympathetic activity is attained by the inhibi-tion of the RVLM via an inhibitory input from the vagal nuclei. See Sections 4.3.2.3, 4.4.1.1, 4.4.1.2, 4.4.2.3 and 4.6.3, for details.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

678 Journal of Psychopharmacology 27(8)

(Adidharma et al., 2012). As both the LH/PF (see Section 5.3.2) and the DR (see Section 5.4.1.3) send excitatory projec-tions to the LC, the LH/PF→DR→LC route is likely to aug-ment the activation of LC via the SC→DMH→LC route, and thus enhance the wakefulness-promoting effect of light. Light is expected to stimulate the sleep-promoting neurones of the VLPO in both diurnal and nocturnal animals: however, in diurnal animals, the wakefulness-promoting effect deriving from activation of the LC is likely to supersede the sleep-pro-moting effect arising from the activation of the VLPO.

7. Neuropathology of the LC

The functional significance of the central noradrenergic system, and in particular of the LC, is highlighted by pathological condi-tions that damage the system. The pathological conditions include chronic progressive neurodegenerative processes and acute (mainly vascular) insults.

The neuromelanin-containing catecholaminergic neurones of the brainstem, including the noradrenergic neurones of the LC, are susceptible to degeneration in progressive neurodegenerative

Figure 9. Schematic diagram of the neuronal network responsible for mediating the acute effects of light on arousal and sympathetic activity (‘photomodulation‘).Adrenoceptors: α1 (excitatory), α2 (inhibitory), β1 (excitatory).Cellular groups (nuclei and ganglia): CG: ciliary ganglion; DMH: dorsomedial hypothalamus; EWN: Edinger-Westphal nucleus; IML: intermedial lateral column; LC: locus coeruleus; OPT: olivary pretectal nucleus; PVN: paraventricular nucleus; SCG: superior cervical ganglion; SCN: suprachiasmatic nucleus; StG: stellate ganglion; TMN: tuber-omammillary nucleus; VLPO: ventrolateral preoptic area.Neurotransmitters: ACh: acetylcholine; GABA: γ-aminobutyric acid; Glu: glutamate; H: histamine; NA: noradrenaline; Ox: orexin; VP: vasopressin. Outputs from nuclei and ganglia: arrows (red: excitatory; blue: inhibitory). Light, either directly or indirectly, via rods and cones, stimulates intrinsically photosensitive retinal ganglion cells (ipRGCs) in the retina, that project to the SCN and VLPO (retino-hypothalamic tract) and OPT (light reflex pathway). Light has a dual effect on arousal. By stimulating sleep-promoting GABAergic neurones in the VLPO, it is sleep-promoting. Light, however, also has a wakefulness-promoting effect via activation of glutamatergic neurones in the SCN which in turn stimulate, via the DMH, the wakefulness-promoting neurones in the LC. The sleep-promoting effect of light seems to predominate in nocturnal animals, which, however, is superseded by the wakefulness-promoting effect in diurnal animals. Light also exerts a dual effect on sympathetic activity. Light can inhibit sympathetic outflow by stimulating GABAergic inhibitory neurones in the SCN, which in turn inhibit premotor sympathetic neurones in the PVN. The withdrawal of the stimulatory influence of the PVN on preganglionic sympathetic neurones in the IML leads to sympathoinhibition in the spinal segments projecting to the superior cervical ganglion, manifesting as pupil constriction and suppression of melatonin synthesis. Light, however, can also stimulate sympathetic activity at lower thoracic levels of the spinal cord, leading to increased cardiac activ-ity. The LC is involved in mediating this effect. The stimulation of glutamatergic excitatory neurones in the SCN by light leads to the activation of the LC, via the DMH. The LC, in turn, can activate preganglionic sympathetic neurones in the IML projecting to the stellate ganglion, leading to cardiostimulation. See Section 6.3, for details.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 679

disorders, such as the synucleinopathies (e.g. Parkinson‘s disease (PD), Lewy body dementia) (Puschmann et al., 2012) and the tau-pathies (e.g. Alzheimer‘s disease (AD)) (Lee et al., 2001).

It has been shown that neurodegenerative changes in PD ascend from the lower brainstem to the cerebral cortex (Braak et al., 2003). It has also been shown that the noradrenergic neu-rones in the LC get involved earlier than the dopaminergic neu-rones in the midbrain, the degeneration of which is responsible for the characteristic motor symptoms of PD (Del Tredici et al., 2002). There is evidence of loss of noradrenergic neurones in the LC in the brains of PD patients, as shown by both post-mortem examination (Bertrand et al., 1997; Zarow et al., 2003) and meas-urement of the neuromelanin signal in vivo by nuclear magnetic resonance (Sasaki et al., 2006).

Progressive degeneration of the LC in PD is likely to be responsible for a number of non-motor symptoms of PD which may appear before the onset of the motor symptoms. The loss of the noradrenergic input to the neocortex, especially to the prefron-tal cortex, has been implicated in early cognitive changes in PD (Vazey and Aston-Jones, 2012) and the development of dementia in more advanced disease (Chan-Palay, 1991; Zweig, 1993). Deficient noradrenergic function may underlie, at least partly, the sleep disorders observed in PD, such as excessive daytime sleepi-ness (Arnulf et al., 2002) and REM sleep behavior disorder (Barber and Dashtipour, 2012; Dugger et al., 2012). The loss of the noradrenergic input to the limbic system has been implicated in the emotional changes, such as depression, often observed in PD (Remy et al., 2005).

There is also a loss of noradrenergic neurones from the LC in another neurodegenerative disease, AD (Busch et al., 1997; Lyness et al., 2003; Tomlinson et al., 1981; Yang et al., 1999; Zarow et al., 2003). The degeneration of LC neurones in AD has been implicated in cognitive changes in this disorder, both at an early stage (Grudzien et al., 2007) and also later as manifest dementia supervenes (Bondareff et al., 1987; Chan-Palay, 1991; Matthews et al., 2002). Depression is a common concomitant of AD, and the depression appears to be related to neuronal loss in the LC (Chan-Palay and Asan, 1989; Forstl et al., 1994).

There are some physiological changes observed in patients suffering from AD that can be attributed to compromised LC function. There is evidence of reduced sympathetic activity and alertness in AD patients (for review, see Samuels and Szabadi, 2008b) consistent with the roles of the LC in the regulation of autonomic activity (see Section 6.2) and arousal (see Section 6.1). Both smooth pursuit and saccadic eye movements have been reported to be impaired in AD patients (for review, see Samuels and Szabadi, 2008b): this may reflect a deficit of the noradrener-gic regulation of motoneurones in the oculomotor complex (see Section 4.4.3.1).

Loss of noradrenergic neurones from the LC has also been found in some other neurodegenerative diseases, such as Down‘s syndrome (German et al., 1992; Lockrow et al., 2011, 2012; Mann et al., 1983), Pick‘s disease (Arima and Akashi, 1990; Takauchi et al., 1995), progressive supranu-clear palsy (Arima et al., 1992; Mann et al., 1983; Wang et al., 2006) and corticobasal degeneration (Gibb et al., 1989).

Recently it has been reported that the LC is damaged in mul-tiple sclerosis (Polak et al., 2011). The damage of the LC in mul-tiple sclerosis may contribute to the clinical picture: it has been

suggested that compromised LC function may be responsible for the deficit of arousal in this condition, manifesting as fatigue (Niepel et al., 2013). Furthermore, the identification of a noradr-energic deficit in multiple sclerosis may open new therapeutic avenues (McSharry, 2011).

Acute, mainly vascular, lesions can also compromise central noradrenergic function. It has been reported that a patient with a focal lesion in the dorsomedial pontine tegmentum, involving the LC, developed both narcolepsy and REM sleep behavior dis-order (Mathis et al., 2007) and this highlights the importance of the LC in mediating the influence of the orexinergic system on wakefulness and muscle tone. It has been shown that brainstem stroke leads to coma only if a well-defined area of the upper pontine tegmentum, including the LC, is affected (Parvizi and Damasio, 2003). This observation suggests that the coma may be due to destruction of the LC, and is consistent with the role of the LC in the maintenance of wakefulness. However, it has been reported recently that damage to some other structures in the vicinity of the LC (parabrachial nucleus, precoeruleus area) may play a major role in the causation of brainstem coma (Fuller et al., 2011).

8. Pathophysiology of the LCThe activity of the LC is altered in a number of pathological and drug-induced states when changed LC activity contributes to the clinical manifestation of these conditions.

The LC plays an important role in both the processing and modulation of pain. The LC is an important relay nucleus in the pain-processing system: it receives nociceptive inputs from both the sensory trigeminal nuclei and the dorsal horn of the spinal cord (Craig, 1992). The LC projects to higher pain-processing centers, such as the somatosensory thalamus (Voisin et al., 2005) and the somatosensory cerebral cortex (Levitt et al., 1984). There is extensive evidence that noxious stimulation leads to an increase in LC activity (for review see Szabadi, 2012). The LC not only receives an input from nociceptive afferents but it also projects to primary pain-sensitive/pain-processing sensory neurones in the trigeminal sensory nuclei of the brainstem (Tsuruoka et al., 2003), the dorsal horn of the spinal cord (Liu et al., 2007), and secondary pain-processing neurones in the somatosensory thalamus (Voisin et al., 2005). Through these projections the LC exerts an inhibi-tory influence on pain sensation (for reviews see Ossipov et al., 2010; Pertovaara and Almeida, 2006; Szabadi, 2012; Willis and Westlund, 1997).

Anxiety is generally defined as an emotional state evoked by threatening stimuli (Szabadi and Bradshaw, 1988), when the threat is direct or immediate, and fear as anticipation of threat (Grillon et al., 1991). The LC is closely involved with both the generation of the emotional states of fear and anxiety and the pro-duction of the behavioral syndrome characteristic of anxious states (Millan, 2003). The basis of this dual role of the LC in fear and anxiety is its two-way relationship to the amygdala, the struc-ture that is critical for the generation of anxiety and anxious responses (Canteras et al., 2010; Charney et al., 1998; Damasio, 1998; LeDoux, 1998; Pare and Duvarci, 2012). As discussed above, the LC sends an excitatory projection to the central and basal nuclei of the amygdala (see Section 4.2.1) and the same nuclei of the amygdala receive an excitatory input from the LC (see Section 5.2). Stimulation of the LC can evoke anxiety

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

680 Journal of Psychopharmacology 27(8)

(McDougle et al., 1995; Redmond and Huang, 1979) and lesions of the central noradrenergic system lead to attenuation of fear-related responses (Redmond, 1987; Verleye and Bernet, 1983). On the other hand, there is evidence that anxious states lead to the activation of the LC (Charney et al., 1990; Cullinan et al., 1995; Rasmussen and Jacobs, 1986).

The effect of threat (anticipatory anxiety) has been studied extensively using the paradigm of ‘fear conditioning‘. Fear condi-tioning involves the association of a noxious (unconditioned) stimulus with a neutral (conditioned) stimulus, leading to the abil-ity of the conditioned stimulus to evoke the response to the uncon-ditioned stimulus (usually pain). Fear conditioning leads to the activation of both the amygdala (Davis et al., 1993; Pascoe and Kapp, 1985) and the LC (Ishida et al., 2002; Rasmussen and Jacobs, 1986). It has been shown that fear-conditioning can modu-late physiological reflexes: the acoustic startle response, involv-ing the contraction of striated muscles, is potentiated whereas the pupillary light reflex, involving the contraction of smooth muscles in the iris, is inhibited by conditioned fear (for review see Szabadi, 2012). In both cases, the activation of the LC, via the amygdala, is involved: increased LC activity enhances the startle response via facilitating motoneurone activity (see Sections 4.4.3 and 4.6.2) and attenuates the pupillary light reflex via enhancing the noradr-energic inhibition of the parasympathetic output to the iris (see Section 6.2) (Szabadi, 2012).

Post-traumatic stress disorder (PTSD) is a clinical syndrome that arises following exposure to extreme stressful experience. Increased startle reactivity is a feature of the syndrome, and it may be related to increased LC activity (Szabadi, 2012). It has been reported that exposure to stress leads to the activation of the LC (Pacák and Palkovits, 2001; Valentino and Van Bockstaele, 2008), and there is evidence of enhanced noradrenergic activity in PTSD (Jacobsen et al., 2001; Southwick et al., 1993).

The LC has been implicated in anesthesia, due to its position in the LC ↔ VLPO → TMN → cerebral cortex circuit (see Section 6.1 and Figure 6) (Nelson et al., 2002, 2003). GABAergic anes-thetics (e.g. propofol, pentobarbitone) stimulate GABAA recep-tors on the GABAergic neurones of the VLPO: this leads to increased inhibition of the wakefulness-promoting histaminergic neurones of the TMN. The withdrawal of the histaminergic stimu-lation of the cerebral cortex results in sedation (Nelson et al., 2002). It should be noted, however, that the central role of the GABAergic inhibition of the TMN in the mode of action of GABAergic anesthetics has been challenged recently (Zecharia et al., 2012).

The LC, by inhibiting the VLPO via stimulation of α2-adrenoceptors, is expected to attenuate the effect of GABAergic anesthetics. This is demonstrated by the effects of drugs that inter-act with inhibitory α2-adrenoceptors (autoreceptors) on LC neu-rones, and thereby modulate LC activity. Thus α2-adrenoceptor agonists (e.g. clonidine, dexmedetomidine) potentiate or even induce anesthesia (Kushikata et al., 2002; Nelson et al., 2003) whereas α2-adrenoceptor antagonists (e.g. yohimbine) antagonize anesthesia (Kushikata et al., 2002). Consistent with these find-ings, potentiation of anesthesia was reported following destruc-tion of the noradrenergic output from the LC by the neurotoxin DSP-4 (Kushikata et al, 2011). It should be noted that the inhibi-tory link between the VLPO and the LC operates in both direc-tions (see Sections 5.3.1 and 6.1). Therefore GABAergic anesthetics, by increasing the activity of VLPO neurones, dampen LC activity (Hirota and Kushikata, 2001).

Rett syndrome (RTT) is a neurodevelopmental disorder caused by loss-of-functions mutations in the methyl-CpG-binding pro-tein-2 (MECP2) gene (Amir et al., 1999). The clinical phenotype is characterized by cognitive, motor and behavioral deficits, ste-reotype hand movements, loss of speech, autistic features, sei-zures, impairments of respiratory, cardiac and gastrointestinal functions (Chahrour and Zoghbi, 2007; Hagberg et al., 1983). Reduced levels of noradrenaline metabolites have been detected in the cerebrospinal fluid of RTT patients (Zoghbi et al., 1989) raising the possibility of a central noradrenergic abnormality. This possibility has been followed up using an animal model of the disorder (MECP2 null mutant mouse). A number of abnormalities of the central noradrenergic system have been detected, including some specifically related to the LC. The abnormalities detected include reduced noradrenaline content (Samaco et al., 2009; Viemari et al., 2005), smaller noradrenergic neurones (Taneja et al., 2009), reduced activity of the synthesizing enzymes tyros-ine hydroxylase (Roux et al., 2010; Samaco et al., 2009; Taneja et al., 2009; Viemari et al., 2005; Zhang et al., 2010) and dopa-mine β-hydroxylase (Zhang et al., 2010), and electrical hyperex-citability of LC neurones (Taneja et al., 2009). Some of the clinical features of the syndrome have been related to central noradrener-gic deficit, such as disordered respiratory control (Viemari et al., 2005; Voituron et al., 2010) and autism (Mehler and Purpura, 2009).

9. ConclusionsThe central noradrenergic neurone has retained the basic charac-teristics of the peripheral sympathetic neurone. The terminal axons arborize extensively and contain both synaptic (‘bouton ter-minaux‘) and non-synaptic sites of release. The non-synaptic release of noradrenaline complements the neurotransmitter action of noradrenaline with a local hormone-like (‘paracrine‘) function. The central noradrenergic neurone has the same pharmacologi-cally-sensitive sites as the peripheral sympathetic neurones: fir-ing- and release-modulating adrenoceptors, sites for the reuptake and re-storage of noradrenaline (cellular and vesicular noradrena-line transporters), machinery of storage, synthesis and degrada-tion. Furthermore, the same types of adrenoceptor are activated by noradrenaline in the periphery and the CNS, and the same cotrans-mitters occur in both peripheral and central noradrenergic neurones.

The central neurone, however, also has some unique features. Central noradrenergic neurones aggregate in distinct nuclei in the brainstem from where they send multiple outputs to diverse struc-tures of the neuraxis. There is evidence that one single neurone may innervate different, and often quite remote structures. Furthermore, the central noradrenergic neurone is a ‘dual function neurone‘: it can exert both excitatory effects, via the activation of α1-adrenoceptors, and inhibitory effects, via the activation of α2-adrenoceptors, on follower cells. In this way the central noradren-ergic neurone does not conform to Eccles‘ principle which states that ‘in the mammalian CNS neurones are either excitatory or inhibitory and are never ambivalent‘ (see page 229 in Eccles, 1976). The duality of its function equips the central noradrenergic neurone with unique flexibility and targeted executive control that can be attained by single function neurones only via the involve-ment of interneurones. The duality of central noradrenergic func-tion is illustrated by the sleep/arousal and autonomic regulatory networks: noradrenergic neurones of the LC stimulate

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 681

wakefulness-promoting and inhibit sleep-promoting mechanisms, and they stimulate sympathetic and inhibit parasympathetic activity.

Of the central noradrenergic nuclei, the LC is the largest and best studied. The multiple and diverse efferent and afferent con-nections of the LC indicate its involvement in the regulation of a large number of physiological functions. However, the impor-tance of central noradrenergic control is best illustrated by consid-ering some integrated physiological functions in which the LC plays a major role. The LC plays a central role in both the sleep/arousal and autonomic networks: via its connections within these networks the LC exerts a powerful wakefulness-promoting and sympathetic-activating influence. Via its position in these net-works, the LC provides an important link between arousal and autonomic activity, fulfilling the role of the ‘arousal/autonomic activity interface‘. The LC also plays a unique role in the network responsible for mediating the acute effect of light on physiological functions, such as arousal and sympathetic activity (‘photomodu-lation‘). In this network, the LC is responsible for mediating the stimulatory effects of light on arousal and sympathetic activity which oppose the inhibitory effects of light on the same functions mediated via a different route (VLPO, PVN). As the inhibitory effects of light appear to be predominant in nocturnal animals, whereas the stimulatory effects prevail in diurnal animals, LC activity seems to be necessary for the daytime functioning of diur-nal animals.

Understanding the functional neuroanatomy of the central noradrenergic system helps to interpret the effects of physiologi-cal variables (e.g. pain, anxiety and stress) and pharmacological agents (e.g. stimulants, sedatives and antidepressants) on the brain and behavior (Samuels and Szabadi, 2008a). As the central noradrenergic system, and in particular the LC, is susceptible to early and progressive degeneration in neurodegenerative diseases (e.g. PD, AD), some clinical features of these diseases (e.g. cogni-tive decline, sleep disorder and depression) can be understood on the basis of the loss of the noradrenergic control of brain areas implicated in the symptoms.

AcknowledgementsThe author is grateful to Rob Langley for drawing the figures.

Conflict of interestThe author declares that there is no conflict of interest.

FundingThis research received no specific grant from any funding agency in the public, commercial, or non-profit sectors.

ReferencesAcosta-Martinez M, Fiber JM, Brown RD, et al. (1999) Localization of

α1B-adrenergic receptor in female rat brain regions involved in stress and neuroendocrine function. Neurochem Int 35: 383–391.

Adidharma W, Leach G and Yan L (2012) Orexinergic signaling mediates light-induced neuronal activation in the dorsal raphe nucleus. Neuro-science 220: 201–207.

Aldes LD (1990) Topographically organized projections from the nucleus subcoeruleus to the hypoglossal nucleus in the rat: A light and elec-tron microscopic study with complementary axonal transport tech-niques. J Comp Neurol 302: 643–656.

Aldes LD, Shaw B, Chronister RB, et al. (1990) Catecholamine-contain-ing axon terminals in the hypoglossal nucleus of the rat: An immuno-electronmicroscopic study. Exp Brain Res 81: 167–178.

Allen GV and Cechetto DF (1994) Serotoninergic and nonserotoninergic neurons in the medullary raphe system have axon collateral projec-tions to autonomic and somatic cell groups in the medulla and spinal cord. J Comp Neurol 350: 357–366.

Altimus CM, Güler AD, Villa KL, et al. (2008) Rods-cones and melanop-sin detect light and dark to modulated sleep independent of image formation. Proc Nat Acad Sci USA 105: 19998–20003.

Amaral DG and Sinnamon HM (1977) The locus coeruleus: Neurobiology of a central noradrenergic nucleus. Prog Neurobiol 9: 147–196.

Amir RE, Van den Veyver IB, Wan M, et al. (1999) Rett syndrome is caused by mutations in X-linked MECP2, encoding methyl-CpG-binding protein 2. Nat Genet 23: 185–188.

Andrews GD and Lavin A (2006) Methylphenidate increases cortical excitability via activation of alpha-2 noradrenergic receptors. Neuro-psychopharmacology 31: 594–601.

Appel NM and Elde RP (1988) The intermediolateral cell column of the thoracic spinal cord is comprised of target-specific subnuclei: Evi-dence from retrograde transport studies and immunohistochemistry. J Neurosci 8: 1767–1775.

Arima K and Akashi T (1990) Involvement of the locus coeruleus in Pick‘s disease with or without Pick body formation. Acta Neuro-pathol 79: 629–633.

Arima K, Murayama S, Oyanagi S, et al. (1992) Presenile dementia with progressive supranuclear palsy tangles and Pick bodies: An unusual degenerative disorder involving the cerebral cortex, cerebral nuclei, and brain stem nuclei. Acta Neuropathol 84: 128–134.

Arnsten AF and Goldman-Rakic PS (1984) Selective prefrontal cortical projections to the region of the locus coeruleus and raphe nuclei in rhesus monkey. Brain Res 306: 9–18.

Arnulf I, Konofal E, Merino-Andreu M, et al. (2002) Parkinson‘s disease and sleepiness. Neurology 58: 1019–1024.

Astier B, Van Bockstaele EJ, Aston-Jones G, et al. (1990) Anatomical evidence for multiple pathways leading from the rostral ventrolateral medulla (nucleus paragigantocellularis) to the locus coeruleus in the rat. Neurosci Lett 118: 141–146.

Aston-Jones G (2005) Brain structures and receptors involved in alertness. Sleep Med 6: S3–S7.

Aston-Jones G and Bloom FE (1981) Activity of norepinephrine-contain-ing locus coeruleus neurons in behaving rats anticipates fluctuations in the sleep-waking cycle. J Neurosci 1: 876–886.

Aston-Jones G and Cohen JD (2005) An integrative theory of locus coe-ruleus-norepinephrine function: Adaptive gain and optimal perfor-mance. Ann Rev Neurosci 28: 403–450.

Aston-Jones G, Astier B and Ennis M (1992) Inhibition of noradrenergic locus coeruleus neurons by C1 adrenergic cells in the rostral ventral medulla. Neuroscience 48: 371–381.

Aston-Jones G, Chen S, Zhu Y, et al. (2001) A neural circuit for circadian regulation of arousal. Nature Neurosci 4: 732–738.

Aston-Jones G, Chiang C and Alexinsky T (1991a) Discharge of norad-renergic locus coeruleus neurons in behaving rats and monkeys sug-gests a role in vigilance. Prog Brain Res 88: 501–520.

Aston-Jones G, Ennis M, Pieribone VA, et al. (1986) The brain nucleus locus coeruleus: Restricted afferent control of a broad efferent net-work. Science 234: 734–737.

Aston-Jones G, Shipley MT, Chouvet G, et al. (1991b) Afferent regulation of locus coeruleus neurons: Anatomy, physiology and pharmacology. Prog Brain Res 88: 47–75.

Bajic D and Proudfit HK (1999) Projections of neurons in the periaqueduc-tal gray to pontine and medullary catecholamine cell groups involved in the modulation of nociception. J Comp Neurol 405: 359–379.

Bajic D, Proudfit HK and Van Bockstaele EJ (2000) Periaqueductal gray neurons monosynaptically innervate extranuclear dendrites in the rat pericoerulear region. J Comp Neurol 427: 649–662.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

682 Journal of Psychopharmacology 27(8)

Bajic D, Van Bockstaele and Proudfit HK (2001) Ultrastructural analysis of ventrolateral periaqueductal gray projections to the A7 catechol-amine cell group. Neuroscience 104: 181–197.

Bajic D, Van Bockstaele and Proudfit HK (2012) Ultrastructural analy-sis of rat ventrolateral periaqueductal gray projections to the A5 cell group. Neuroscience 224: 145–159.

Bakes A, Bradshaw CM and Szabadi E (1990) Attenuation of the pupillary light reflex in anxious patients. Br J Clin Pharmacol 30: 377–381.

Baldo BA, Daniel RA, Berridge CW, et al. (2003) Overlapping distribu-tions of orexin/hypocretin- and dopamine-beta-hydroxylase immuno-reactive fibers in rat brain regions mediating arousal, motivation, and stress. J Comp Neurol 464: 220–237.

Barber A and Dashtipour K (2012) Sleep disturbances in Parkinson‘s dis-ease with emphasis on rapid eye movement sleep behaviour disorder. Int J Neurosci 122: 407–412.

Barman SM and Gebber GL (1985) Axonal projection patterns of ven-trolateral medullospinal sympathoexcitatory neurons. J Neurophysiol 53: 1551–1566.

Bateman RJ, Boychuk CR, Philbin KE, et al. (2012) β-adrenergic recep-tor modulation of neurotransmission to cardiac vagal neurons in the nucleus ambiguous. Neuroscience 210: 58–66.

Baumann N and Pham-Dinh D (2001) Biology of oligodendrocyte and myelin in the mammalian central nervous system. Physiol Rev 81: 871–927.

Baver SB, Pickard GE, Sollars PJ, et al. (2008) Two types of melanopsin retinal ganglion cell differentially innervate the hypothalamic supra-chiasmatic nucleus and the olivary pretectal nucleus. Eur J Neurosci 27: 1763–1770.

Bay KD, Mamiya K, Good CH, et al. (2006) Alpha-2 adrenergic regula-tion of pedunculopontine nucleus neurons during development. Neu-roscience 141: 769–779.

Beaudet A and Descarries L (1978) The monoamine innervation of the rat cerebral cortex: Synaptic and nonsynaptic axon terminals. Neurosci-ence 3: 851–860.

Beckstead RM, Domesick VB and Nauta WJ (1979) Efferent connections of the substantia nigra and ventral tegmental area in the rat. Brain Res 175: 191–217.

Ben-Jonathan N and Hnasko R (2001) Dopamine as a prolactin (PRL) inhibitor. Endocr Rev 22: 724–763.

Berridge CW and Abercrombie ED (1999) Relationship between locus coeruleus discharge rates and rates of norepinephrine release within neocortex as assessed by in vivo microdialysis. Neuroscience 93: 1263–1270.

Berridge CW and Foote SL (1991) Effects of locus coeruleus activation on electroencephalographic activity in neocortex and hippocampus. J Neurosci 11: 3135–3145.

Berridge CW, Isaac SO and España RA (2003) Additive wake-promoting actions of medial basal forebrain noradrenergic α1- and β-receptor stimulation. Behav Neurosci 117: 350–359.

Berridge CW, Page ME, Valentino RJ, et al. (1993) Effects of locus coeru-leus inactivation on electroencephalographic activity in neocortex and hippocampus. Neuroscience 55: 381–393.

Bertrand E, Lechowicz W, Szpak GM, et al. (1997) Qualitative and quan-titative analysis of locus coeruleus neurons in Parkinson‘s disease. Folia Neuropathol 35: 80–86.

Bie B, Fields HL, Williams JT, et al. (2003) Roles of α1- and α2-adrenoceptors in the nucleus raphe magnus in opioid analgesia and opioid abstinence-induced hyperalgesia. J Neurosci 23: 7950–7957.

Bienkowski MS and Rinaman L (2008) Noradrenergic inputs to the paraventricular hypothalamus contribute to hypothalamic-pituitary-adrenal axis and central Fos activation in rats after acute systemic endotoxin exposure. Neuroscience 156: 1093–1102.

Bitsios P, Szabadi E and Bradshaw CM (1996) The inhibition of the light reflex by the threat of an electric shock: A potential laboratory model of human anxiety. J Psychopharmacol 10: 279–287.

Blake TJ, Tillery CE and Reynolds GP (1998) Antipsychotic drug affini-ties at alpha2-adrenoceptor subtypes in post-mortem human brain. J Psychopharmacol 12: 151–154.

Bogerts B (1981) A brainstem atlas of catecholaminergic neurons in man, using melanin as a natural marker. J Comp Neurol 197: 63–80.

Bondareff W, Mountjoy CQ, Roth M, et al. (1987) Neuronal degenera-tion in locus coeruleus and cortical correlates of Alzheimer disease. Alzheimer Dis Assoc Disord 1: 256–262.

Booze RM, Crisostomo EA and Davis JN (1989) Species differences in the localization and number of CNS beta adrenergic receptors: Rat versus guinea pig. J Pharmacol Exp Ther 249: 911–920.

Boundy VA and Cincotta AH (2000) Hypothalamic adrenergic receptor changes in the metabolic syndrome of genetically obese (ob/ob) mice. Am J Physiol Regul Integr Comp Physiol 279: R505–R514.

Bourgin P, Huitrón-Reséndiz S, Spier AD, et al. (2000) Hypocretin-1 modulates rapid eye movement sleep through activation of locus coe-ruleus neurons. J Neurosci, 20: 7760–7765.

Braak H, Del Tredici K, Rüb U, et al. (2003) Staging of brain pathol-ogy related to sporadic Parkinson‘s disease. Neurobiol Aging 24: 197–211.

Breen LA, Burde R and Loewy AD (1983) Brainstem connections to the Edinger-Westphal nucleus of the cat: A retrograde tracer study. Brain Res 261: 303–306.

Brown RE, Sergeeva OA, Eriksson KS, et al. (2002) Convergent excita-tion of dorsal raphe serotonin neurons by multiple arousal systems (orexin/hypocretin, histamine and noradrenaline). J Neurosci 22: 8850–8859.

Bubar MJ and Cunningham KA (2007) Distribution of serotonin 5-HT2C receptors in the ventral tegmental area. Neuroscience 146: 286–297.

Buffalari DM and Grace AA (2007) Noradrenergic modulation of baso-lateral amygdala neuronal activity: Opposing influences of α-2 and β receptor activation. J Neurosci 27: 12358–12366.

Buijs RM, la Fleur SE, Wortel J, et al. (2003) The suprachiasmatic nucleus balances sympathetic and parasympathetic output to periph-eral organs through separate preautonomic neurons. J Comp Neurol 464: 36–48.

Burnstock G (2004) Cotransmission. Curr Opin Pharmacol 4: 47–52.Burnstock G (2009) Purinergic cotransmission. Exp Physiol 94: 20–24.Busch C, Bohl J and Ohm TG (1997) Spatial, temporal and numeric analy-

sis of Alzheimer changes in the nucleus coeruleus. Neurobiol Aging 18: 401–406.

Byrum CE and Guyenet PG (1987) Afferent and efferent connections of the A5 noradrenergic cell group in the rat. J Comp Neurol 261: 529–542.

Cahusac PMB, Morris R and Hill RG (1995) A pharmacological study of the modulation of neuronal and behavioural nociceptive responses in the rat trigeminal region. Brain Res 700: 70–82.

Cajochen C, Münch M, Kobialka S, et al. (2005) High sensitivity of human melatonin, alertness, thermoregulation, and heart rate to short wavelength light. J Clin Endocrin Metab 90: 1311–1316.

Cameron AA, Khan IA, Westlund KN, et al. (1995) The efferent projec-tions of the periaqueductal gray in the rat: A Phaseolus vulgaris-leu-coagglutinin study. II Descending projections. J Comp Neurol 351: 585–601.

Campbell G (1987) Cotransmission. Ann Rev Pharmacol Toxicol 27: 51–70.Canteras NS, Resstel LB, Bertoglio LJ, et al. (2010) Neuroanatomy of

anxiety. Curr Top Behav Neurosci 2: 77–96.Carpenter MB, Periera AB and Guha N (1992) Immunocytochemistry

of oculomotor afferents in the squirrel monkey (Saimiri sciureus). J Hirnforsch 33: 151–167.

Catus SL, Gibbs ME, Sato M, et al. (2011) Role of β-adrenoceptors in glucose uptake in astrocytes using β-adrenoceptor knockout mice. Br J Pharmacol 162: 1700–1715.

Cechetto DF and Saper CB (1990) Role of the cerebral cortex in auto-nomic function. In: Loewy AD and Spyer KM (eds) Central Regula-tion of Autonomic Functions. Oxford: Oxford University Press, pp. 208–223.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 683

Cedarbaum JM and Aghajanian GK (1978) Afferent projections to the rat locus coeruleus as determined by a retrograde tracing technique. J Comp Neurol 178: 1–16.

Chahrour M and Zoghbi HY (2007) The story of Rett syndrome: From clinic to neurobiology. Neuron 56: 58–65.

Chan-Palay V (1991) Alterations in the locus coeruleus in dementias of Alzheimer‘s and Parkinson‘s disease. Prog Brain Res 88: 625–630.

Chan-Palay V and Asan E (1989) Alterations in catecholamine neurons of the locus coeruleus in senile dementia of the Alzheimer‘s type and in Parkinson‘s disease with and without dementia and depression. J Comp Neurol 287: 373–392.

Charney DS and Redmond DE Jr (1983) Neurobiological mechanisms in human anxiety. Neuropharmacology 22: 1531–1536.

Charney DS, Grillon C and Bremner JD (1998) The neurobiological basis of anxiety and fear: Circuits, mechanisms, and neurochemical interac-tions (Part 1). Neuroscientist 4: 35–44.

Charney DS, Woods SW, Nagy LM, et al. (1990) Noradrenergic function in panic disorder. J Clin Psychiatry 51: S5–S11.

Chen A, Kelley LD and Janušonis S (2012) Effects of prenatal stress and monoaminergic perturbations on the expression of serotonin 5-HT4 and adrenergic β2 receptors in the embryonic mouse telencephalon. Brain Res 1459: 27–34.

Chen F-J and Sara SJ (2007) Locus coeruleus activation by foot shock or electrical stimulation inhibits amygdala neurons. Neuroscience 144: 472–481.

Chen SK, Badea TC and Hattar S (2011) Photoentrainment and pupillary light reflex are mediated by distinct populations of ipRGCs. Nature 476: 92–95.

Chen SY and Chai CY (2002) Coexistence of neurons integrating urinary bladder activity and pelvic nerve activity in the same cardiovascular areas of the pontomedulla in cats. Chin J Physiol 45: 41–50.

Cheng Z, Zhang H, Yu J, et al. (2004) Attenuation of baroreflex sensitiv-ity after domoic acid lesion of the nucleus ambiguus of rats. J Appl Physiol 96: 1137–1145.

Chibuzo GA and Cummings JF (1980) Motor and sensory centers for the innervation of mandibular and sublingual salivary glands: A horserad-ish peroxidase study in the dog. Brain Res 189: 301–313.

Chiti Z and Teschemacher AG (2007) Exocytosis of norepinephrine at axon varicosities and neuronal cell bodies in the rat brain. FASEB J 21: 2540–2550.

Chong W, Li LH, Lee K, et al. (2004) Subtypes of alpha1- and alpha2-adrenoceptors mediating noradrenergic modulation of spontaneous inhibitory postsynaptic currents in the hypothalamic paraventricular nucleus. J Neuroendcrinol 16: 450–457.

Chou TC, Bjorkum AA, Gaus SE, et al. (2002) Afferents to the ventrolat-eral preoptic nucleus. J Neurosci 22: 977–990.

Chronwall BM (1985) Anatomy and physiology of the neuroendocrine arcuate nucleus. Peptides 6: S1–S11.

Civantos Calzada B and Aleixandre de Artiñano A (2001) Alpha-adreno-ceptor subtypes. Pharmacol Res 44: 195–208.

Cohen Z, Molinatti G and Hamel E (1997) Astroglial and vascular interac-tions of noradrenaline terminals in the rat cerebral cortex. J Cerebr Blood Flow Metab 17: 894–904.

Couto B, Moroni CR, dos Reis Ferreira CM, et al. (2006) Descriptive and functional neuroanatomy of locus coeruleus-noradrenaline-contain-ing neurons involvement in bradykinin-induced antinociception on principal sensory trigeminal nucleus. J Chem Neuroanat 32: 28–45.

Craig AD (1992) Spinal and trigeminal lamina I input to the locus coe-ruleus anterogradely labeled with Phaseolus vulgaris leucoagglutinin (PHA-L) in the cat and monkey. Brain Res 584: 325–328.

Crochet S and Sakai K (2003) Dopaminergic modulation of behavioural states in mesopontine tegmentum: A reverse microdialysis study in freely moving cats. Sleep 26: 801–806.

Cullinan WE, Herman JP, Battaglia DF, et al. (1995) Pattern and time course of immediate early gene expression in rat brain following acute stress. Neuroscience 64: 477–505.

Cunningham ET Jr and Sawchenko PE (1988) Anatomical specificity of noradrenergic inputs to the paraventricular and supraoptic nuclei of the rat hypothalamus. J Comp Neurol 274: 60–76.

Daftary SS, Boudaba C and Tasker JG (2000) Noradrenergic regulation of parvocellular neurons in the rat hypothalamic paraventricular nucleus. Neuroscience 96: 743–751.

Dahlström A and Fuxe K (1964) Evidence for the existence of mono-amine-containing neurones in the central nervous system. I. Demon-stration of monoamines in the cell bodies of brain stem neurones. Acta Physiol Scand 62: 232: S1–S55.

Damasio AR (1998) Emotion in the perspective of an integrated nervous system. Brain Res Rev 26: 83–86.

Dampney RA, Furlong TM, Horiuchi J, et al. (2013) Role of dorsolateral preiaqueductal grey in the coordinated regulation of cardiovascular and respiratory function. Auton Neurosci 175: 17–25.

Dampney RAL (1994) Functional organization of central pathways regu-lating the cardiovascular system. Physiol Rev 74: 332–364.

Danysz W, Dyr W, Plaznik A, et al. (1989) The effect of microinjections of clonidine into the locus coeruleus on cortical EEG in rats. Pol J Pharmacol Pharm 41: 45–50.

Date Y, Ueta Y, Yamashita H, et al. (1999) Orexins, orexigenic hypotha-lamic peptides, interact with autonomic, neuroendocrine and neuro-regulatory systems. Proc Natl Acad Sci USA 96: 748–753.

Datta S, O‘Malley MW and Patterson EH (2011) Calcium/calmodulin kinase II in the pedunculopontine tegmental nucleus modulates the initiation and maintenance of wakefulness. J Neurosci 31: 17007–17016.

Davis M, Falls WA, Campeau S, et al. (1993) Fear-potentiated star-tle: A neural and pharmacological analysis. Behav Brain Res 58: 175–198.

Davis H, Guo X, Lambert S, et al. (2012) Small molecule induction of human umbilical stem cells into myelin basic protein positive olo-godendrocytes in defined three-dimensional environment. ACS Chem Neurosci 3: 31–39.

Day HEW, Campeau S, Watson SJ Jr, et al. (1997) Distribution of α1a–, α1b– and α1d–adrenergic receptor mRNA in the rat brain and spinal cord. J Chem Neuroanat 13: 115–139.

Day HEW, Greenwood BN, Hammack SE, et al. (2004) Differential expression of 5HT-1A, α1b adrenergic, CRF-R1, and CRF-R2 recep-tor mRNA in serotonergic, γ-aminobutyric acidergic, and catechol-aminergic cells of the rat dorsal raphe nucleus. J Comp Neurol 474: 364–378.

De Andrés I, Garzόn M and Reinoso-Suárez F (2011) Functional anatomy of non-REM sleep. Front Neurol 2: 1–14.

De Lecea L, Kilduff TS, Peyron C, et al. (1998) The hypocretins: Hypo-thalamus-specific peptides with neuroexcitatory activity. Proc Natl Acad Sci USA 95: 322–327.

Del Cid-Pellitero E and Garzόn M (2011) Hypocretin1/OrexinA-contain-ing axons innervate locus coeruleus neurons that project to the rat medial prefrontal cortex. Implication in the sleep-wakefulness cycle and cortical activation. Synapse 65: 843–857.

Del Tredici K, Rub U, De Vos RA, et al. (2002) Where does Parkinson disease pathology begin in the brain? J Neuropathol Exp Neurol 61: 413–426.

Descarries L and Mechawar N (2000) Ultrastructural evidence for diffuse transmission by monoamine and acetylcholine neurons in the central nervous system. Prog Brain Res 125: 27–47.

Descarries L, Watkins KC and Lapierre Y (1977) Noradrenergic axon terminals in the cerebral cortex of rat. III. Topometric ultrastructural analysis. Brain Res 133: 197–222.

Deutch AY, Goldstein M and Roth RH (1986) Activation of the locus coeruleus induced by selective stimulation of the ventral tegmental area. Brain Res 363: 307–314.

Domyancic AV and Morilak DA (1997) Distribution of α1A adrenergic receptor mRNA in the rat brain visualized by in situ hybridization. J Comp Neurol 386: 358–378.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

684 Journal of Psychopharmacology 27(8)

Drolet G and Gauthier P (1985) Peripheral and central mechanisms of the pressor response elicited by stimulation of the locus coeruleus in the rat. Can J Physiol Pharmacol 63: 599–605.

Drolet G, Van Bockstaele EJ and Aston-Jones G (1992) Robust enkepha-lin innervation of the locus coeruleus from the rostral medulla. J Neu-rosci 12: 3162–3174.

Dugger BN, Murray ME, Boeve BF, et al. (2012) Neuropathological anal-ysis of brainstem cholinergic and catecholaminergic nuclei in relation to rapid eye movement (REM) sleep behaviour disorder. Neuroathol Appl Neurobiol 38: 142–152.

Ebert U (1996) Noradrenalin enhances the activity of cochlear nucleus neurons in the rat. Eur J Neurosci 8: 1306–1314.

Eccles J (1976) From electrical to chemical transmission in the central nervous system. Notes Rec R Soc Lond 30: 219–230.

Egan TM and North RA (1985) Acetylcholine acts on m2-muscarinic receptors to excite rat locus coeruleus neurones. Br J Pharmacol 85: 733–735.

Eisenach JC, De Kock M and Klimscha W (1996) Alpha(2)-adrener-gic agonists for regional anesthesia. A clinical review of clonidine (198401995). Anesthesiology 85: 655–674.

El Mansari M, Sakai K and Jouvet M (1989) Unitary characteristics of presumptive cholinergic tegmental neurons during the sleep-waking cycle in freely moving cats. Exp Brain Res 76: 519–529.

Elam M, Svensson TH and Thorén P (1986) Locus coeruleus neurons and sympathetic nerves: Activation by cutaneous sensory afferents. Brain Res 366: 254–261.

Engberg G and Svensson TH (1980) Pharmacological analysis of a cholin-ergic receptor mediated regulation of brain norepinephrine neurons. J Neural Transm 49: 137–150.

Ennis M and Aston-Jones G (1986) A potent excitatory input to the nucleus locus coeruleus from the ventrolateral medulla. Neurosci Lett 71: 299–305.

Ennis M and Aston-Jones G (1988) Activation of locus coeruleus from nucleus paragigantocellularis: A new excitatory amino acid pathway in brain. J Neurosci 8: 3644–3657.

Ennis M and Aston-Jones G (1989a) GABA-mediated inhibition of locus coeruleus from the dorsomedial rostral medulla. J Neurosci 9: 2973–2981.

Ennis M and Aston-Jones G (1989b) Potent inhibitory input to locus coeru-leus from the nucleus prepositus hypoglossi. Brain Res Bull 22: 793–803.

Ennis M, Behbehani M, Shipley MT, et al. (1991) Projections from the periaqueductal gray to the rostromedial periocular region and the nucleus locus coeruleus. J Comp Neurol 306: 480–494.

Ericson H, Blomqvist A and Köhler C (1989) Brainstem afferents to the tuberomammillary nucleus in the rat brain with special reference to monoaminergic innervation. J Comp Neurol 281: 169–192.

España RA and Berridge CW (2006) Organisation of noradrenergic effer-ents to arousal-related basal forebrain structures. J Comp Neurol 496: 668–683.

España RA, Reis KM, Valentino RJ, et al. (2005) Organization of hypo-cretin/orexin efferents to locus coeruleus and basal forebrain arousal-related structures. J Comp Neurol 481: 160–178.

Fallon JH, Koziell DA and Moore RY (1978) Catecholamine innervation of the basal forebrain. II. Amygdala, suprarhinal cortex and entorhinal cortex. J Comp Neurol 180: 509–532.

Farb CR, Chang W and LeDoux JE (2010) Ultrastructural characterization of noradrenergic axons and beta-adrenergic receptors in the lateral nucleus of the amygdala. Front Behav Neurosci 4: 1–9.

Felten DL and Sladek Jr, JR (1983) Monoamine distribution in primate brain. V. Monoaminergic nuclei: Anatomy, pathways and local orga-nization. Brain Res Bull 10: 171–284.

Flügge G, van Kampen M, Meyer H, et al. (2003) Alpha2A and apha2C-adrenoceptor regulation in the brain: Alpha2A changes persist after chronic stress. Eur J Neurosci 17: 917–928.

Foote SL, Berridge CW, Adams LM, et al. (1991) Electrophysiological evidence for the involvement of the locus coeruleus in alerting, orient-ing, and attending. Prog Brain Res 88: 521–532.

Foote SL, Bloom FE and Aston-Jones G (1983) Nucleus locus coeruleus: New evidence of anatomical and physiological specificity. Physiol Rev 63: 844–914.

Forstl H, Levy R, Burns A, et al. (1994) Disproportionate loss of noradren-ergic and cholinergic neurons as cause of depression in Alzheimer‘s disease - a hypothesis. Pharmacopsychiatry 27: 11–15.

Fort P, Bassetti CL and Luppi PH (2009) Alternating vigilance states: New insights regarding neuronal networks and mechanisms. Eur J Neurosci 29: 1741–1753.

Fort P, Khateb A, Pegna A, et al. (1995) Noradrenergic modulation of cholinergic nucleus basalis neurons demonstrated by in vitro pharma-cological and immunohistochemical evidence in the guinea-pig brain. Eur J Neurosci 7: 1502–1511.

Fortaleza EAT, Scopinho AA and Corrêa FMA (2012) Paraventricu-lar and supraoptic nuclei of the hypothalamus mediate cardiovas-cular responses evoked by the microinjection of noradrenaline into the medial amygdaloid nucleus of the rat brain. Neuroscience 219: 157–165.

Fritschy JM and Grzanna R (1990) Demonstration of two separate descending noradrenergic pathways to the rat spinal cord: Evidence for an intragriseal trajectory of locus coeruleus axons in the superfi-cial layers of the dorsal horn. J Comp Neurol 291: 553–582.

Fu Y, Matta SG, McIntosh JM, et al. (1999) Inhibition of nicotine-induced hippocampal norepinephrine release in rats by alpha-conotoxins MII and AuIB microinjected into the locus coeruleus. Neurosci Lett 266: 113–116.

Fukami H and Bradley RM (2005) Biophysical and morphological proper-ties of parasympathetic neurons controlling the parotid and von Ebner salivary glands in rats. J Neurophysiol 93: 678–686.

Fukuda A, Minami T, Nabekura J, et al. (1987) The effects of noradrena-line on neurones in the rat dorsal motor nucleus of the vagus, in vitro. J Physiol 393: 213–231.

Fuller PM, Gooley JJ and Saper CB (2006) Neurobiology of the sleep-wake cycle: sleep architecture, circadian regulation, and regulatory feedback. J Biol Rhythms 21: 482–493.

Fuller PM, Saper CB and Lu J (2007) The pontine REM switch: Past and present. J Physiol 584: 735–741.

Fuller P, Sherman D, Pedersen NP, et al. (2011) Reassessment of the structural basis of the ascending arousal system. J Comp Neurol 519: 933–956.

Fung SJ, Chan JY, Manzoni D, et al. (1994a) Cotransmitter-mediated locus coeruleus action on motoneurons. Brain Res Bull 35: 423–432.

Fung SJ, Reddy VK, Zhuo H, et al. (1994b) Anatomical evidence for the presence of glutamate or enkephalin in noradrenergic projection neu-rons of the locus coeruleus. Microsc Res Tech 29: 219–225.

Furst S (1999) Transmitters involved in antinociception in the spinal cord. Brain Res Bull 48: 129–141.

Fuxe K, Dahlström JG, Marcellino D, et al. (2010) The discovery of cen-tral monoamine neurons gave volume transmission to the wired brain. Prog Neurobiol 90: 82–100.

Gallopin T, Fort P, Eggermann E, et al. (2000) Identification of sleep-promoting neurons in vitro. Nature 404: 992–995.

Gatter KC and Powell TPS (1977) The projection of the locus coeru-leus upon the neocortex in the macaque monkey. Neuroscience 2: 441–445.

Geffen LB and Livett BG (1971) Synaptic vesicles in sympathetic neu-rons. Physiol Rev 51: 98–157.

German DC, Manaye KF, White CL, III, et al. (1992) Disease-specific patterns of locus coeruleus cell loss. Ann Neurol 32: 667–676.

Gervasoni D, Darracq L, Fort P, et al. (1998) Electrophysiological evi-dence that noradrenergic neurons of the rat locus coeruleus are toni-cally inhibited by GABA during sleep. Eur J Neurosci 10: 964–970.

Ghiasvand M, Rezayof A, Ahmadi S, et al. (2011) β1-noradrenergic sys-tem of the central amygdala is involved in state-dependent memory induced by a cannabinoid agonist, WIN55,212, in rat. Behav Brain Res 225: 1–6.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 685

Gibb WR, Luthert PJ and Marsden CD (1989) Corticobasal degeneration. Brain 112: 1171–1192.

Gilsbach R and Hein L (2012) Are the pharmacology and physiology of α2-adrenoceptors determined by α2-heteroreceptors and autoreceptors respectively? Br J Pharmacol 165: 90–102.

Giuliano F and Rampin O (2000) Central noradrenergic control of penile erection. Int J Impot Res 12: S13–S19.

Glass MJ, Colago EE and Pickel VM (2002) Alpha-2A-adrenergic recep-tors are present on neurons in the central nucleus of the amygdala that project to the dorsal vagal complex in the rat. Synapse 46: 258–268.

Gobert A, Rivet J-M, Lejeune F, et al. (2000) Serotonin2C receptors toni-cally suppress the activity of mesocortical dopaminergic and adren-ergic, but not serotonergic, pathways: A combined dialysis and electrophysiological analysis in the rat. Synapse 36: 205–221.

Gonzalez MMC and Aston-Jones G (2006) Circadian regulation of arousal: Role of noradrenergic locus coeruleus system and light expo-sure. Sleep 29: 1327–1336.

Goodchild AK and Moon EA (2009) Maps of cardiovascular and respira-tory regions of rat ventral medulla: Focus on the caudal medulla. J Chem Neuroanat 38: 209–221.

Gooley JJ, Lu J, Fischer D, et al. (2003) A broad role for melanopsin in nonvisual photoreception. J Neurosci 23: 7093–7106.

Gottesmann. C (2011) The involvement of noradrenaline in rapid eye movement sleep mentation. Front Neurol 2: 1–10.

Goutagny R, Luppi PH, Salvert D, et al. (2008) Role of the dorsal paragi-gantocellular reticular nucleus in paradoxical (rapid eye movement) sleep generation: A combined electrophysiological and anatomical study in the rat. Neuroscience 152: 849–857.

Green GM, Lyons L and Dickenson AH (1998) α2-adrenoceptor antago-nists enhance responses of dorsal horn neurones to formalin induced inflammation. Eur J Pharmacol 347: 201–204.

Greenblatt SH (1999) John Hughlings Jackson and the conceptual founda-tions of the neurosciences. Physis Riv Stor Sci 36: 367–386.

Grillon C, Ameli R, Woods SW, et al. (1991) Fear-potentiated startle in humans: Effects of anticipatory anxiety on the acoustic blink reflex. Psychophysiology 28: 588–595.

Gritti I, Maiville L and Jones BE (1993) Codistribution of GABA- with acetylcholine-synthesizing neurons in the basal forebrain of the rat. J Comp Neurol 329: 438–457.

Grudzien A, Shaw P, Weintraub S, et al. (2007) Locus coeruleus neuro-fibrillary degeneration in aging, mild cognitive impairment and early Alzheimer‘s disease. Neurobiol Aging 28: 327–335.

Guistina A and Veldhuis JD (1998) Pathophysiology of the neuroregu-lation of growth hormone secretion in experimental animals and humans. Endocr Rev 19: 717–797.

Güler AD, Ecker JL, Lall GS, et al. (2008) Melanopsin cells are the princi-pal conduits for rod-cone input to non-image-forming vision. Nature 453, 102–106.

Guo N-N and Li B-M (2007) Cellular and subcellular distributions of β1- and β2-adrenoceptors in the CA1 and CA3 regions of the rat hippo-campus. Neuroscience 146: 298–305.

Gurtu S, Pant KK, Sinha JN, et al. (1984) An investigation into the mecha-nism of cardiovascular responses elicited by electrical stimulation of locus coeruleus and subcoeruleus in the cat. Brain Res 301: 59–64.

Guyenet PG (1991) Central noradrenergic neurons: The autonomic con-nection. Prog Brain Res 88: 365–380.

Guyenet PG and Young BS (1987) Projections of nucleus paragigantocel-lularis lateralis to locus coeruleus and other structures in rat. Brain Res 406: 171–184.

Guyenet PG, Stornetta RL, Riley T, et al. (1994) Alpha 2A-adrenergic receptors are present in lower brainstem catecholaminergic and sero-tonergic neurons innervating spinal cord. Brain Res 638: 285–294.

Haas HL and Lin J-S (2012) Waking with the hypothalamus. Pflugers Arch – Eur J Physiol 463: 31–42.

Haas H and Panula P (2003) The role of histamine and the tuberomammil-lary nucleus in the nervous system. Nature Rev Neurosci 4: 121–130.

Hagan JJ, Leslie RA, Patel S, et al. (1999) Orexin A activates locus coe-ruleus cell firing and increases arousal in the rat. Proc Natl Acad Sci USA 96: 10911–10916.

Hagberg B, Aicardi J, Dias K, et al. (1983) A progressive syndrome of autism, dementia, ataxia, and loss of purposeful hand use in girls: Rett‘s syndrome. Ann Neurol 14: 471–479.

Hamill RW (1996) Peripheral autonomic nervous system. In: Robertson D, Low PA and Polinsky RJ (eds) Primer on the Autonomic Nervous System. San Diego: Academic Press, pp. 12–25.

Hancock MB and Fougerousse CL (1976) Spinal projections from the nucleus locus coeruleus and nucleus subcoeruleus in the cat and mon-key as demonstrated by the retrograde transport of horseradish peroxi-dase. Brain Res Bull 1: 229–234.

Hartman BK, Zide D and Undenfriend S (1972) The use of dopamine β-hydroxylase as a marker for the central noradrenergic nervous sys-tem in the rat brain. Proc Nat Acad Sci USA 69: 2722–2726.

Head GA, Chan CKS and Burke SL (1998) Relationship between imid-azoline and α 2-adrenoceptors involved in the sympatho-inhibitory actions of centrally acting antihypertensive agents. J Auton Nerv Syst 72: 163–169.

Heinricher MM, Taveras I, Leith JL, et al. (2009) Descending control of nociception: Specificity, recruitment and plasticity. Brain Res Rev 60: 214–225.

Herbert H and Saper CB (1992) Organization of medullary adrenergic and noradrenergic projections to the periaqueductal gray matter in the rat. J Comp Neurol 315: 34–52.

Herman JP, Figueiredo H, Mueller NK, et al. (2003) Central mechanisms of stress integration: Hierarchical circuitry controlling hypothalamo-pituitary-adrenocortical responsiveness. Front Neuroendocrin 24: 151–180.

Hermann DM, Luppi P-H, Peyron C, et al. (1997) Afferent projections to the rat nuclei raphe magnus, raphe pallidus and reticularis gigantocel-lularis pars α demonstrated by iontophoretic application of cholera-toxin (subunit b). J Chem Neuroanat 13: 1–21.

Herring N and Paterson DJ (2009) Neuromodulators of peripheral cardiac sympatho-vagal balance. Exp Physiol 94: 46–53.

Hertz L, Chen Y, Gibbs ME, et al. (2004) Astrocytic adrenoceptors: A major target in neurological and psychiatric disorders? Curr Drug Targets CNS Neurol Disord 3: 239–268.

Hieble JP and Ruffolo RR (1995) Subclassification of the β-adrenoceptors. Pharmacol Commmun 6: 183–193.

Higo S, Ito K, Fuchs D, et al. (1990) Anatomical interconnections of the pedunculopontine tegmental nucleus and the nucleus prepositus hypoglossi in the cat. Brain Res 536: 79–85.

Hillman KL, Doze VA and Porter JE (2005) Functional characteriza-tion of the β-adrenergic receptor subtypes expressed by CA1 pyra-midal cells in the rat hippocampus. J Pharmacol Exp Ther 314: 561–567.

Hirata H and Aston-Jones G (1994) A novel long-latency response of locus coeruleus neurons to noxious stimuli: Mediation by peripheral C-fibers. J Neurophysiol 71: 1752–1761.

Hirota K and Kushikata T (2001) Central noradrenergic neurones and the mechanism of general anaesthesia. Br J Anaesth 87: 811–812.

Holstege G (1987) Some anatomical observations on the projections from the hypothalamus to brainstem and spinal cord: An HRP and autora-diographic tracing study in the cat. J Comp Neurol 260: 98–126.

Horvath TL, Peyron C, Diano S, et al. (1999) Hypocretin (orexin) acti-vation and synaptic innervation of the locus coeruleus noradrenergic system. J Comp Neurol 415: 145–159.

Hosoya Y, Sugiura Y, Okado N, et al. (1991) Descending input from the hypothalamic paraventricular nucleus to sympathetic preganglionic neurons in the rat. Exp Brain Res 85: 10–20.

Hou RH, Freeman C, Langley RW, et al. (2005) Does modafinil activate the locus coeruleus in man? Comparison of modafinil and clonidine on arousal and autonomic functions in human volunteers. Psycho-pharmacology 181: 537–549.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

686 Journal of Psychopharmacology 27(8)

Hou RH, Samuels ER, Langley RW, et al. (2007) Arousal and the pupil: Why diazepam-induced sedation is not accompanied by miosis. Psy-chopharmacology 195: 41–59.

Hou RH, Scaife J, Freeman C, et al. (2006) Relationship between sedation and pupillary function: Comparison of diazepam and diphenhydr-amine. Br J Clin Pharmacol 61: 752–760.

Hou YP, Manns ID and Jones BE (2002) Immunostaining of cholinergic pontomesencephalic neurons for α1 versus α2 adrenergic receptors suggests different sleep-wake state activities and roles. Neuroscience 114: 517–521.

Hrabovszky E and Liposits Z (1994) Adrenergic innervation of dopamine neurons in the hypothalamic arcuate nucleus of the rat. Neurosci Lett 182: 143–146.

Huang H-P, Zhu F-P, Chen X-W, et al. (2012) Physiology of quantal nor-epinephrine release from somatodendritic sites of neurones in locus coeruleus. Front Mol Neurosci 5: 1–5.

Illes P and Norenberg W (1990) Blockade of alpha 2-adrenoceptors increases opioid mu-receptor-mediated inhibition of the firing rate of rat locus coeruleus neurones. Naunyn-Schmiedebergs Arch Pharma-col 342: 490–496.

Ishida Y, Hashiguchi H, Takeda R, et al. (2002) Conditioned-fear stress increases Fos expression in monoaminergic and GABAergic neurons in the locus coeruleus and dorsal raphe nuclei. Synapse 45: 46–51.

Ishikawa M and Tanaka C (1977) Morphological organization of cate-cholamine terminals in the diencephalon of the rhesus monkey. Brain Res 119: 43–55.

Ishikawa M, Shimada S and Tanaka C (1975) Histochemical mapping of catecholamine neurons and fibre pathways in the pontine tegmentum of the dog. Brain Res 86: 1–6.

Iversen LL (1967) The Uptake and Storage of Noradrenaline in Sympa-thetic Nerves. Cambridge: Cambridge University Press.

Iwase M, Homma I, Shioda S, et al. (1993) Histamine immunoreac-tive neurons in the brain stem of the rabbit. Brain Res Bull 32: 267–272.

Jacobowitz DM and MacLean PD (1978) A brainstem atlas of catechol-aminergic neurons and serotonergic perikarya in a pygmy primate (Cebuella pygmaca). J Comp Neurol 177: 397–416.

Jacobsen LK, Southwick SM and Kosten TR (2001) Substance use dis-orders in patients with posttraumatic stress disorder: A review of the literature. Am J Psychiatry 158: 1184–1190.

Jansen AS, Ter Horst GJ, Mettenleiter TC, et al. (1992) CNS cell groups projecting to the submandibular parasympathetic preganglionic neu-rons in the rat: A retrograde transneuronal viral cell body labeling study. Brain Res 572: 253–260.

Jiménez-Rivera CA, Figueroa J, Vázquez-Torres R, et al. (2012) Pre-synaptic inhibition by α2 receptors in the VTA. Eur J Neurosci 35: 1406–1415.

Jodo E and Aston-Jones G (1997) Activation of locus coeruleus by pre-frontal cortex is mediated by excitatory amino acid inputs. Brain Res 768: 327–332.

Jodo E, Chiang C and Aston-Jones G (1998) Potent excitatory influence of prefrontal cortex activity on noradrenergic locus coeruleus neurons. Neuroscience 83: 3–79.

Johnson AD, Peoples J, Stornetta RL, et al. (2002) Opioid circuits origi-nating from the nucleus paragigantocellularis and their potential role in opiate withdrawal. Brain Res 955: 72–84.

Jones BE (1990) Immunohistochemical study of choline acetyltransfer-ase-immunoreactive processes and cells innervating the pontomedul-lary reticular formation in the rat. J Comp Neurol 295: 485–514.

Jones BE (1991) Paradoxical sleep and its chemical/structural substrates in the brain. Neuroscience 40: 637–656.

Jones BE (2004) Activity, modulation and role of basal forebrain cholin-ergic neurons innervating the cerebral cortex. Prog Brain Res 145: 157–169.

Jones BE (2005) From waking to sleeping: Neuronal and chemical sub-strates. Trends Pharmacol Sci 26: 578–586.

Jones BE and Moore RY (1977) Ascending projections of the locus coe-ruleus in the rat. II Autoradiographic study. Brain Res 127: 25–53.

Jones BE and Yang T-Z (1985) The efferent projections from the reticular formation and the locus coeruleus studied by anterograde and retro-grade axonal transport in the rat. J Comp Neurol 242: 56–92.

Jones LS, Gauger LL and Davis JN (1985) Anatomy of brain alpha1-adrenergic receptors: In vitro autoradiography with [125I]-HEAT. J Comp Neurol 231: 190–208.

Joyce JN, Lexow N, Soo JK, et al. (1992) Distribution of beta-adrenergic receptor subtypes in human post-mortem brain: Alterations in limbic regions of schizophrenics. Synapse 10: 228–246.

Kalsbeek A, Cutrera RA, Van Heerikhulze JJ, et al. (1999) GABA release from suprachiasmatic nucleus is necessary for the light-induced inhi-bition of nocturnal melatonin release in the rat. Neuroscience 91: 453–461.

Kalsbeek A, Garidou ML, Palm IF, et al. (2000) Melatonin sees the light: Blocking GABA-ergic transmission in the paraventricular nucleus induces daytime secretion of melatonin. Eur J Neurosci 12: 3146–3154.

Kalsbeek A, Palm IF, La Fleur SE, et al. (2006) SCN outputs and the hypothalamic balance of life. J Biol Rhythms 21: 458–469.

Kang Y-M, Ouyang W, Chen J-Y, et al. (2000) Norepinephrine modulates single hypothalamic arcuate neurons via α1 and β adrenergic recep-tors. Brain Res 869: 146–157.

Kapoor V, Minson J and Chalmers J (1992) Ventral medulla stimulation increases blood pressure and spinal cord amino acid release. Neuro-report 3: 55–58.

Kasparov S and Teschemacher AG (2008) Altered central catechol-aminergic transmission and cardiovascular disease. Exp Physiol 93: 725–740.

Kaur S, Saxena RN and Mallick BN (2001) GABAergic neurons in prep-ositus hypoglossi regulate REM sleep by its action on locus coeruleus in freely moving rats. Synapse 42: 141–150.

Kawasaki A and Kardon RH (2007) Intrinsically photosensitive retinal ganglion cells. J Neuro-Ophthalmol 27: 195–204.

Kayama Y, Ohta M and Jodo E (1992) Firing of ‘possibly‘ cholinergic neurons in the rat laterodorsal tegmental nucleus during sleep and wakefulness. Brain Res 569: 210–220.

Kiernan JA (2005) Barr‘s The Human Nervous System: An Anatomical Viewpoint, 8th ed. Philadelphia: Lippincott Williams & Wilkins.

Kim M, Chiego DJ and Bradley RM (2004a) Morphology of parasympa-thetic neurons innervating rat lingual salivary glands. Auton Neurosci 111: 27–36.

Kim M-A, Lee HS, Lee BY, et al. (2004b) Reciprocal connections between subdivisions of the dorsal raphe and the nuclear core of the locus coeruleus in the rat. Brain Res 1026: 56–67.

Kimura F and Nakamura S (1985) Locus coeruleus neurons in the neona-tal rat: Electrical activity and responses to sensory stimulation. Brain Res 355; 301–305.

Klepper A and Herbert H (1991) Distribution and origin of noradrenergic and serotonergic fibers in the cochlear nucleus and inferior colliculus of the rat. Brain Res 557: 190–201.

Kobayashi H, Frattola L, Ferrarese C, et al. (1982) Characterization of beta-adrenergic receptors on human cerebral microvessels. Neurology 32: 1384–1387.

Korotkova TM, Sergeeva OA, Ponomarenko AA, et al.(2005) Histamine excites noradrenergic neurons in locus coeruleus in rats. Neurophar-macology 49: 129–134.

Koss MC (1986) Pupillary dilation as an index of central nervous system α2-adrenoceptor activation. J Pharmacol Met 15: 1–19.

Koss MC, Gherezghiher T and Nomura A (1984) CNS adrenergic inhi-bition of parasympathetic oculomotor tone. J Auton Nerv Syst 10: 55–68.

Kromer LF and Moore RY (1980) Norepinephrine innervation of the cochlear nuclei by locus coeruleus neurons in the rat. Anat Embryol (Berl) 158: 227–244.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 687

Kumagai H, Oshima N, Matsuura T, et al. (2012) Importance of rostral ventrolateral medulla neurons in determining efferent sympathetic nerve activity and blood pressure. Hypertens Res 35: 132–141.

Kupfermann I (1991) Functional studies of cotransmission. Physiol Rev 71: 683–732.

Kushikata T, Hirota K, Yoshida H, et al. (2002) Alpha-2 adrenocep-tor activity affects propofol-induced sleep time. Anesth Analg 94: 1201–1206.

Kushikata T, Yoshida H, Kudo M, et al. (2011) Role of coerulean norad-renergic neurones in general anaesthesia in rats. Br J Anaesth 107: 924–929.

Kwiat GC and Basbaum AI (1992) The origin of brainstem noradrenergic and serotonergic projections to the spinal cord dorsal horn in the rat. Somatosens Mot Res 9: 157–173.

Larkman PM and Kelly JS (1992) Ionic mechanisms mediating 5-hydroxy-tryptamine- and noradrenaline-evoked depolarization of adult rat facial motoneurones. J Physiol 456: 473–490.

LeDoux J (1998) Fear and the brain: Where have we been, and where are we going? Biol Psychiatry 44: 1229–1238.

Lee HS, Kim M-A and Waterhouse BD (2005a) 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 481: 179–193.

Lee HS, Lee BY and Waterhouse BD (2005b) Retrograde study of projec-tions from the tuberomammillary nucleus to the dorsal raphe and the locus coeruleus in the rat. Brain Res 1043: 65–75.

Lee VM, Goedert M and Trojanowski JQ (2001) Neurodegenerative tau-pathies. Annu Rev Neurosci 24: 1121–1159.

Le Maître E, Barde SS, Palkovits M, et al. (2013) Distinct features of neurotransmitter systems in the human brain with focus on the galanin system in locus coeruleus and dorsal raphe. Proc Natl Acad Sci USA 110: E536–545.

Leong SK, Shieh JY and Wong WC (1984) Localizing spinal-cord-pro-jecting neurons in adult albino rats. J Comp Neurol 228: 1–17.

LePard KJ, Ren J and Galligan JJ (2004) Presynaptic modulation of cholinergic and non-cholinergic fast synaptic transmission in the myenteric plexus of guinea pig ileum. Neurogastroenterol Motil 16: 355–364.

Levitt P and Moore RY (1979) Origin and organization of brainstem cat-echolamine innervation in the rat. J Comp Neurol 186: 505–528.

Levitt P, Rakic P and Goldman-Rakic P (1984) Region-specific distribu-tion of catecholamine afferents in primate cerebral cortex: A fluores-cence histochemical analysis. J Comp Neurol 227: 23–36.

Lewis DI and Coote JH (1990) Excitation and inhibition of rat sympathetic preganglionic neurones by catecholamines. Brain Res 530: 229–234.

Li D-P, Atnip LM, Chen S-R, et al. (2005) Regulation of synaptic inputs to paraventricular-spinal output neurons by α2 adrenergic receptors. J Neurophysiol 93: 393–402.

Li H-Y, Ericsson A and Sawchenko PE (1996) Distinct mechanisms underlie activation of hypothalamic neurosecretory neurons and their medullary catecholaminergic afferents in categorically different stress paradigms. Proc Natl Acad Sci USA 93: 2359–2364.

Li X, Rainnie DG, McCarley RW, et al. (1998) Presynaptic nicotinic receptors facilitate monoaminergic transmission. J Neurosci 18: 1904–1912.

Li Y and Van den Pol AN (2005) Direct and indirect inhibition by cate-cholamines of hypocretin/orexin neurons. J Neurosci 25: 173–183.

Lin J-S, Anaclet C, Sergeeva OA, et al. (2011) The waking brain: An update. Cell Mol Life Sci 68: 2499–2512.

Liu L, Tsuruoka M, Maeda M, et al. (2007) Coerulospinal inhibition of visceral nociceptive processing in the rat spinal cord. Neurosi Lett 426: 139–144.

Livett BG (1973) Histochemical visualization of peripheral and central adrenergic neurones. Br Med Bull 29: 93–99.

Lockley SW, Evans EE, Scheer FA, et al. (2006) Short-wavelength sensitivity for the direct effects of light on alertness, vigilance, and the waking electroencephalogram in humans. Sleep 29: 161–168.

Lockrow J, Boger G, Aston-Jones G, et al. (2011) A noradrenergic lesion exacerbates neurodegeneration in a Down syndrome mouse model. J Alzheimer Dis 23: 471–489.

Lockrow JP, Fortress AM and Granholm A-CE (2012) Age-related neu-rodegeneration and memory loss in Down syndrome. Curr Gerontol Geriatr Res 2012. doi: 10.1155/2012/463909.

Loewi O (1921) Über humorale Uebertragbarkeit der Herzenwirkung. Pflügers Arch ges Physiol 237: 504–514.

Loewy AD (1981) Raphe pallidus and raphe obscurus projections to the intermediolateral cell column in the rat. Brain Res 222: 129–133.

Loewy AD (1990a) Anatomy of the autonomic nervous system: An over-view. In: Loewy AD and Spyer KM (eds) Central Regulation of Auto-nomic Functions. Oxford: Oxford University Press, pp. 3–16.

Loewy AD (1990b) Central autonomic pathways. In: Loewy AD and Spyer KM (eds) Central Regulation of Autonomic Functions. Oxford: Oxford University Press, pp. 88–103.

Loewy AD and Spyer KM (1990) Vagal preganglionic neurons. In: Loewy AD and Spyer KM (eds) Central Regulation of Autonomic Functions. Oxford: Oxford University Press, pp. 68–87.

Loewy AD, McKellar S and Saper CB (1979) Direct projections from A5 catecholamine cell group to the intermedio-lateral cell column. Brain Res 174: 309–314.

Loughlin SE, Foote SL and Bloom FE (1986a) Efferent projections of nucleus locus coeruleus: Topographic organization of cells of origin demonstrated by three-dimensional reconstruction. Neuroscience 18: 291–306.

Loughlin SE, Foote SL and Fallon JH (1982) Locus coeruleus projections to the cortex: Topography, morphology and collaterization. Brain Res Bull 9: 287–294.

Loughlin SE, Foote SL and Grzanna R (1986b) Efferent projections of nucleus locus coeruleus: Morphologic subpopulations have different efferent targets. Neuroscience 18: 307–319.

Lu J, Jhou TC and Saper CB (2006) Identification of wake-active dopami-nergic neurons in the ventral periaqueductal gray matter. J Neurosci 26: 193–202.

Lu J, Shiromani P and Saper CB (1999) Retinal input to the sleep-active ventrolateral preoptic nucleus in the rat. Neuroscience 93: 209–214.

Luiten PGM, Ter Horst GJ, Karst H, et al (1985) The course of the para-ventricular hypothalamic efferents to autonomic structures in medulla and spinal cord. Brain Res 329: 374–378.

Lung MA (1994) Mechanisms of sympathetic enhancement and inhibition of parasympathetically induced salivary secretion in anaesthetized dogs. Br J Pharmacol 112: 411–416.

Luppi P-H, Aston-Jones G, Akaoka H, et al. (1995) Afferent projections to the rat locus coeruleus demonstrated by retrograde and anterograde tracing with cholera-toxin B subunit and Phaseolus vulgaris leucoag-glutinin. Neuroscience 65: 119–160.

Luppi PH, Clement O, Sapin E, et al. (2012) Brainstem mechanisms of paradoxical (REM) sleep generation. Pflugers Arch 463: 43–52.

Lyness SA, Zarow C and Chui HC (2003) Neuron loss in key cholinergic and aminergic nuclei in Alzheimer disease: A meta-analysis. Neuro-biol Aging 24: 1–23.

Lyons DJ, Hellysaz R and Broberger C (2012) Prolactin regulates tuber-oinfundibular dopamine neuron discharge pattern: Novel feedback control mechanisms in the lactotrophic axis. J Neurosci 32: 8074–8083.

Lyons WE and Grzanna R (1988) Noradrenergic neurons with diver-gent projections to the motor trigeminal nucleus and the spinal cord: A double retrograde neuronal labeling study. Neuroscience 26: 681–693.

McBride RL and Sutin J (1976) Projections of the locus coeruleus and adjacent pontine tegmentum in the cat. J Comp Neurol 165: 265–284.

McCormick DA, Pape HC and Williamson A (1991) Actions of norepi-nephrine in the cerebral cortex and thalamus: Implications for func-tion of the central noradrenergic system. Prog Brain Res 88: 293–305.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

688 Journal of Psychopharmacology 27(8)

McDougle CJ, Krystal JH, Price LH, et al. (1995) Noradrenergic response to acute ethanol administration in healthy subjects: Comparison with intravenous yohimbine. Psychopharmacology 118: 127–135.

McGinty DJ and Harper RM (1976) Dorsal raphe neurons: Depression of firing during sleep in cats. Brain Res 101: 569–575.

Machado BH and Brody MJ (1990) Mechanisms of pressor response produced by stimulation of nucleus ambiguus. Am J Physiol 259: R955–R962.

McKitrick DJ and Calaresu FR (1996) Nucleus ambiguus inhibits activity of cardiovascular units in RVLM. Brain Res 742 203–210.

McSharry C (2011) Multiple sclerosis: Noradrenaline deficits suggest novel MS treatments. Nat Rev Neurol 7: 184.

Mann DM, Yates PO and Hawkes J (1983) The pathology of the human locus coeruleus. Clin Neuropathol 2: 1–7.

Manns ID, Alonso A and Jones BE (2000a) Discharge properties of juxta-cellularly labelled and immunohistochemically identified cholinergic basal forebrain neurons recorded in association with the electroen-cephalogram in anesthetized rats. J Neurosci 20: 1505–1518.

Manns ID, Alonso A and Jones BE (2000b) Discharge profiles of juxtacel-lularly labelled and immunohistochemically identified GABAergic basal forebrain neurons recorded in association with the electroen-cephalogram in anesthetized rats. J Neurosci 20: 9252–9263.

Manns ID, Lee MG, Modirrousta M, et al. (2003) Alpha 2 adrenergic receptors on GABAergic, putative sleep-promoting basal forebrain neurons. Eur J Neurosci 18: 723–727.

Margalit D and Segal M (1979) A pharmacologic study of analgesia pro-duced by stimulation of the nucleus locus coeruleus. Psychopharma-cology 62: 169–173.

Marotte LR and Raisman G (1974) An experimental investigation of the relationship between granular vesicles and noradrenergic axons in the preoptic area of the rat. Brain Res 70: 335–339.

Martinez-Peña y, Valenzuela I, Rogers RC, Hermann E, et al.(2004) Norepinephrine effects on identified neurons of the rat dorsal motor nucleus of the vagus. Am J Physiol Gastrointest Liver Physiol 286: G333–G339.

Mathis J, Hess CW and Bassetti C (2007) 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 78: 427–429.

Matsutani K, Tsuruoka M, Shinya A, et al. (2000) Stimulation of the locus coeruleus suppresses trigeminal sensorimotor function in the rat. Brain Res Bull 53: 827–832.

Matthews KL, Chen CPL-H, Esiri MM, et al. (2002) Noradrenergic changes, aggressive behavior, and cognition in patients with demen-tia. Biol Psychiatry 51: 407–416.

Mehler MF and Purpura DP (2009) Autism, fever, epigenetics and the locus coeruleus. Brain Res Rev 59: 388–392.

Michimori A, Araki K and Hagiwara H (1997) Effects of illuminance lev-els on sympathetic activities. Sleep Res 26: 677.

Millan MJ (2002) Descending control of pain. Prog Neurobiol 66: 355–474.

Millan MJ (2003) The neurobiology and control of anxious states. Prog Neurobiol 70: 83–244.

Millan MJ, Lejeune F and Gobert A (2000) Reciprocal autoreceptor and heteroreceptor control of serotonergic, dopaminergic and noradren-ergic transmission in the frontal cortex: Relevance to the actions of antidepressant agents. J Psychopharmacol 14: 114–138.

Mistlberger RE (2005) Circadian regulation of sleep in mammals: Role of the suprachiasmatic nucleus. Brain Res Rev 49: 429–454.

Modirrousta M, Mainville L and Jones BE (2004) GABAergic neurons with α2-adrenergic receptors in basal forebrain and preoptic area express c-Fos during sleep. Neuroscience 129: 803–810.

Monti JM (2011) Serotonin control of sleep-wake behaviour. Sleep Med Rev 15: 269–281.

Monti JM and Jantos H (2008) The roles of dopamine and serotonin, and their receptors, in regulating sleep and waking. Prog Brain Res 172: 625–646.

Monti JM and Monti D (2007) The involvement of dopamine in the modu-lation of sleep and waking. Sleep Med Rev 11: 113–133.

Moore RY and Bloom FE (1979) Central catecholamine neuron systems: Anatomy and physiology of the norepinephrine and epinephrine sys-tems. Ann Rev Neurosci 2: 113–168.

Morrison JH, Foote SL, O‘Connor D, et al. (1982) Laminar, tangential and regional organization of the noradrenergic innervation of monkey cortex: Dopamine-β-hydroxylase immunohistochemistry. Brain Res Bull 9: 309–319.

Morrison JH, Molliver ME, Grzanna R, et al. (1979) Noradrenergic inner-vation patterns in three regions of medial cortex: An immunofluores-cence characterization. Brain Res Bull 4: 849–857.

Morrison SF and Nakamura K (2011) Central neural pathways for thermo-regulation. Front Biosci 16: 74–104.

Mosqueda-Garcia R (1996) Central autonomic regulation. In: Robertson D, Low PA and Polinsky RJ (eds) Primer on the Autonomic Nervous System. San Diego: Academic Press, pp. 3–12.

Nagai T, Satoh K, Imamoto K, et al. (1981) Divergent projections of cat-echolamine neurons of the locus coeruleus as revealed by fluorescent retrograde double labelling technique. Neurosci Lett 23: 117–123.

Nakamura K, Matsumura K, Hübschle T, et al. (2004) Identification of sympathetic premotor neurons in medullary raphe regions mediat-ing fever and other thermoregulatory functions. J Neurosci 24: 5370–5380.

Nakamura K, Matsumura K, Kobayashi S, et al. (2005) Sympathetic pre-motor neurons mediating thermoregulatory function. Neurosci Res 51: 1–8.

Nelson LE, Guo TZ, Lu J, et al. (2002) The sedative component of anes-thesia is mediated by GABAA receptors in an endogenous sleep path-way. Nature Neuroscience 5: 979–984.

Nelson LE, Lu J, Guo TZ, et al. (2003) The α2-adrenoceptor agonist dex-medetomidine converges on an endogenous sleep-promoting pathway to exert its sedative effects. Anesthesiology 98: 428–436.

Nemoto T, Konno A and Chiba T (1995) Synaptic contact of neuropep-tide-and amine-containing axons on parasympathetic preganglionic neurons in the superior salivatory nucleus of the rat. Brain Res 685: 33–45.

Nicholson JE and Severin CM (1981) The superior and inferior salivatory nuclei in the rat. Neurosci Lett 21: 149–154.

Nicola SM, Taha SA, Kim SW, et al. (2005) Nucleus accumbens dopa-mine release is necessary and sufficient to promote the behav-ioural response to reward- predictive cues. Neuroscience 135; 1025–1033.

Niepel G, Bibani RH, Vilisaar J, et al. (2013) Association of a deficit of arousal with fatigue in multiple sclerosis: Effect of modafinil. Neuro-pharmacology 664: 380–388.

Nieuwenhuys R (1985) Chemoarchitecture of the Brain. Berlin: Springer-Verlag.

Nishino H, Koizumi K and Brooks CM (1976) The role of suprachias-matic nuclei of the hypothalamus in the production of circadian rhythm. Brain Res 112: 45–59.

Nitz D and Siegel JM (1997) GABA release in the locus coeruleus as a function of sleep/wake state. Neuroscience 78: 795–801.

Nosaka S, Yasunaga K and Tamai S (1982) Vagal cardiac preganglionic neurons: Distribution, cell types, and reflex discharges. Am J Physiol 243: R92–R98.

Nygren L-G and Olson L (1977) 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 132: 85–93.

Olave MJ and Maxwell DJ (2002) An investigation of neurones that pos-sess the α2C-adrenergic receptor in the rat dorsal horn. Neuroscience 115: 31–40.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 689

Olson L and Fuxe K (1971) On the projections from the locus coeruleus noradrenaline neurons: The cerebellar innervation. Brain Res 28: 165–171.

O‘Malley SS, Chen TB, Francis BE, et al. (1998) Characterization of specific binding of [125I]L-762,459, a selective alpha1A-adreno-ceptor radioligand to rat and human tissues. Eur J Pharmacol 348: 287–295.

Ornstein K, Milon H, McRae-Degueurce A, et al. 1987) Biochemical and radioautographic evidence for dopaminergic afferents of the locus coeruleus originating in the ventral tegmental area. J Neural Transm 70: 183–191.

Osaka T and Matsumura H (1994) Noradrenergic inputs to sleep-related neurons in the preoptic area from the locus coeruleus and the ventro-lateral medulla in the rat. Neurosci Res 19: 39–50.

Ossipov MH, Dussor GO and Porreca F (2010) Central modulation of pain. J Clin Invest 120: 3779–3787.

Owesson CA, Seif I, McLaughlin DP, et al. (2003) Different alpha(2) adrenoceptor subtypes control noradrenaline release and cell firing in the locus coeruleus of wildtype and monoamine oxidase-A knockout mice. Eur J Neurosci 18: 34–42.

Pacák K and Palkovits M (2001) Stressor specificity of central neuroen-docrine responses: Implications for stress-related disorders. Endocr Rev 22: 502–548.

Pal D and Mallick BN (2007) Neural mechanism of rapid eye movement sleep generation with reference to REM-OFF neurons in locus coeru-leus. Indian J Med Res 125: 721–739.

Palacios JM, Hoyer D and Cortés R (1987) α1-Adrenoceptors in the mam-malian brain: Similar pharmacology but different distribution in rodents and primates. Brain Res 419: 65–75.

Palkovits M, Baffi JS and Dvori S (1995) Neuronal organization of stress response. Pain-induced c-fos expression in brain stem catecholamin-ergic cell groups. Ann NY Acad Sci 29; 313–326.

Palkovits M, Záborsky L, Feminger A, et al. (1980) Noradrenergic inner-vation of the rat hypothalamus: Experimental biochemical and elec-tron microscopic studies. Brain Res 191: 161–171.

Pan ZZ, Grudt TJ and Williams JT (1994) Alpha 1-adrenoceptors in rat dorsal raphe neurons: Regulation of two potassium conductances. J Physiol (Lond) 478: 437–447.

Papadopoulos GC and Parnavelas JG (1991) Monoamine systems in the cerebral cortex: Evidence for anatomical specificity. Prog Neurobiol 36: 195–200.

Papay R, Gaivin R, Jha A, et al. (2006) 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 497: 209–222.

Pare D and Duvarci S (2012) Amygdala microcircuits mediating fear expression and extinction. Curr Opin Neurobiol 22: 717–723.

Parkis MA, Bayliss DA and Berger AJ (1995) Actions of norepinephrine on rat hypoglossal motoneurons. J Neuropsysiol 74: 1911–1919.

Parvizi J and Damasio AR (2003) Neurochemical correlates of brainstem coma. Brain 126: 1524–1536.

Paschalis A, Churchill L, Marina N, et al. (2009) β1-Adrenoceptor dis-tribution in the rat brain: An immunohistochemical study. Neurosci Lett 458: 84–88.

Pascoe JP and Kapp BS (1985) Electrophysiological characteristics of amygdaloid central nucleus neurons during Pavlovian fear condition-ing in the rabbit. Behav Brain Res 16: 117–133.

Pascual J, del Arco C, Gonzalez AM, et al.(1992) Quantitative light microscope autoradiographic localization of alpha 2-adrenoceptors in the human brain. Brain Res 585: 116–127.

Paspalas CD and Papadopoulos GC (1996) Ultrastructural relationships between noradrenergic nerve fibers and non-neuronal elements in the rat cerebral cortex. Glia 17: 133–146.

Pasquier DA, Kemper TL, Forbes WB, et al. (1977) Dorsal raphe, substan-tia nigra and locus coeruleus: Interconnections with each other and the neostriatum. Brain Res Bull 2: 323–339.

Passatore M (1976) Physiological characterization of efferent cervical sympathetic fibers influenced by changes of illumination. Exp Neurol 53: 71–81.

Passatore M and Pettorossi VE (1976) Efferent fibers in the cervical sym-pathetic nerve influenced by light. Exp Neurol 52: 66–82.

Peever J (2011) Control of motoneuron function and muscle tone during REM sleep, REM sleep behaviour disorder and cataplexy/narcolepsy. Arch Ital Biol 149: 454–466.

Pepper CM and Henderson G (1980) Opiates and opioid peptides hyper-polarize locus coeruleus neurons in vitro. Science 209: 394–395.

Perreau-Lenz S, Kalsbeek A, Garidou ML, et al. (2003) Suprachiasmatic control of melatonin synthesis in rats: Inhibitory and stimulatory mechanisms. Eur J Neurosci 17: 221–228.

Pertovaara A and Almeida A (2006) Endogenous pain modulation: Descending inhibitory systems. In: Cervero F and Jensen TS (eds) Handbook of Clinical Neurology: Pain, vol. 81. Amsterdam: Else-vier, pp. 179–192.

Peyron C, Luppi P-H, Rampon C, et al. (1995) Localization of the GABA-ergic neurons projecting to the rat dorsal raphe and locus coeruleus. Soc Neurosc Abs 21: 373.

Phillips MA, Szabadi E and Bradshaw CM (2000a) Comparison of the effects of clonidine and yohimbine on pupillary diameter at different illumination levels. Br J Clin Pharmacol 50: 65–68.

Phillips MA, Szabadi E and Bradshaw CM (2000b) Comparison of the effects of clonidine and yohimbine on spontaneous pupillary fluctua-tions in healthy human volunteers. Psychopharmacology 150: 85–89.

Pickel VM, Segal M and Bloom FE (1974) A radioautographic study of the efferent pathways of the nucleus locus coeruleus. J Comp Neurol 155: 15–42.

Pieribone VA, Nicholas AP, Dagerlind Å, et al. (1994) Distribution of α1 adrenoceptors in rat brain revealed by in situ hybridization experi-ments utilizing subtype-specific probes. J Neurosci 14: 4252–4268.

Poitras D and Parent A (1978) Atlas of the distribution of monoamine-containing nerve cell bodies in the brain stem of the cat. J Comp Neu-rol 179: 699–718.

Polak PE, Kalinin S and Feinstein DL (2011) Locus coeruleus damage and noradrenaline reductions in multiple sclerosis and experimental autoimmune encephalomyelitis. Brain 134: 665–677.

Proudfit HK and Clark FM (1991) The projections of locus coeruleus neu-rons to the spinal cord. Prog Brain Res 88: 123–141.

Pudovkina OL, Cremers TIFH and Westerink BHC (2003) Regulation of the release of serotonin in the dorsal raphe nucleus by α1 and α2 adre-noceptors. Synapse 50: 77–82.

Puschmann A, Bhidayasiri R and Weiner WJ (2012) Synucleinopathies from bench to bedside. Parkinsonism Relat Disord 18: S24–S27.

Pyner S (2009) Neurochemistry of the paraventricular nucleus of the hypothalamus: Implications for cardiovascular regulation. J Chem Neuroanat 38: 197–208.

Qu LL, Guo NN and Li BM (2008) Beta1- and beta2-adrenoceptors in basolateral nucleus of amygdala and their roles in consolidation of fear memory in rats. Hippocampus 18: 1131–1139.

Raichle ME, Hartman BK, Eichling JO, et al. (1975) Central noradrener-gic regulation of cerebral blood flow and vascular permeability. Proc Nat Acad Sci USA 72: 3726–3730.

Rajkowski J, Kubiak P and Aston-Jones G (1993) Correlations between locus coeruleus (LC) neural activity, pupil diameter and behaviour in monkey support a role of LC in attention. Abstr Soc Neurosci 19: 974.

Rajkowski J, Kubiak P and Aston-Jones G (1994) Locus coeruleus activ-ity in monkey: Phasic and tonic changes are associated with altered vigilance. Brain Res Bull 35: 607–616.

Ranson RN, Motawei K, Pyner S, et al. (1998) The paraventricular nucleus of the hypothalamus sends efferents to the spinal cord of the rat that closely appose sympathetic preganglionic neurones projecting to the stellate ganglion. Exp Brain Res 120: 164–172.

Ranson RN, Santer RM and Watson AH (2006) The relationship between serotonin, dopamine beta hydroxylase and GABA immunoreactive

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

690 Journal of Psychopharmacology 27(8)

inputs and spinal preganglionic neurones projecting to the major pel-vic ganglion of Wistar rats. Neuroscience 141: 1935–1949.

Rasmussen K and Aghajanian GK (1990) Serotonin excitation of facial motoneurons: Receptor subtype characterization. Synapse 5: 324–332.

Rasmussen K and Jacobs BL (1986) Single unit activity of locus coeruleus neurons in the freely moving cat. II. Conditioning and pharmacologic studies. Brain Res 23: 335–344.

Rasmussen K, Morilak D and Jacobs BL (1986) Single unit activity of locus coeruleus neurons in the freely moving cat. I. During natural-istic behaviors and in response to simple and complex stimuli. Brain Res 371: 324–334.

Rathner JA, Owens NC and McAllen RM (2001) Cold-activated raphe-spinal neurons in rats. J Physiol 535: 841–854.

Redmond DE (1987) Studies of the nucleus locus coeruleus in monkeys and hypotheses for neuropsychopharmacology. In: Meltzer HY (ed) Psychopharmacology: The Third Generation of Progress. New York: Raven Press, pp. 467–974.

Redmond DE and Huang YH (1979) New evidence for a locus coeruleus-norepinephrine connection with anxiety. Life Sci 25: 2149–2162.

Remy P, Doder M, Lees A, et al. (2005) Depression in Parkinson‘s dis-ease: Loss of dopamine and noradrenaline innervation in the limbic system. Brain 128: 1314–1322.

Revell VL, Arendt J, Fogg LF, et al. (2006) Alerting effects of light are sensitive to very short wavelengths. Neurosci Lett 399: 96–100.

Reyes BAS, Valentino RJ, Xu G, et al. (2005) Hypothalamic projections to locus coeruleus neurons in rat brain. Eur J Neurosci 22: 93–106.

Rivot JP, Chiang CY and Besson JM (1982) Increase of serotonin metabo-lism within the dorsal horn of the spinal cord during nucleus raphe magnus stimulation, as revealed by in vivo electrochemical detection. Brain Res 238: 117–126.

Robertson HA and Leslie RA (1985) Noradrenergic alpha 2 binding sites in vagal dorsal motor nucleus and nucleus tractus solitarius: Autora-diographic localization. Can Physiol Pharmacol 63: 1190–1194.

Rodriguez M, Carillon C, Coquerel A, et al. (1995) Evidence for the pres-ence of beta 3-adrenergic receptor mRNA in the human brain. Brain Res Mol Brain Res 29: 369–375.

Roux JC, Panyotis N, Dura E, et al. (2010) Progressive noradrenergic defi-cits in the locus coeruleus of Mecp2 deficient mice. J Neurosci Res 88: 1500–1509.

Saigal RP, Karamanlidis A, Voogd J, et al. (1980) 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 197: 200–206.

Saint-Mleux B, Eggermann E, Bisetti A, et al. (2004) Nicotinic enhance-ment of the noradrenergic inhibition of sleep-promoting neurons in the ventrolateral preoptic area. J Neurosci 24: 63–67.

Sakai K (1988) Executive mechanisms of paradoxical sleep. Arch Ital Biol 126: 239–257.

Sakai K, Salvert D, Touret M, et al. (1977) Afferent connections of the nucleus raphe dorsalis in the cat as visualized by the horseradish per-oxidase technique. Brain Res 137: 11–35.

Samaco RC, Mandel-Brehm C, Chao HT, et al. (2009) Loss of MeCP2 in aminergic neurons causes cell-autonomous defects in neurotransmit-ter synthesis and specific behavioural abnormalities. Proc Natl Acad Sci USA 106: 21966–21971.

Samuels ER and Szabadi E (2008a) Functional neuroanatomy of the nor-adrenergic locus coeruleus: Its roles in the regulation of arousal and autonomic function. Part I: Principles of functional organisation. Curr Neuropharmacol 6: 235–253.

Samuels ER and Szabadi E (2008b) Functional neuroanatomy of the noradrenergic locus coeruleus: Its roles in the regulation of arousal and autonomic function. Part II: Physiological and pharmacological manipulations and pathological alterations of locus coeruleus activity in humans. Curr Neuropharmacol 6: 254–285.

Samuels ER, Hou RH, Langley RW, et al. (2006) Comparison of prami-pexole and modafinil on arousal, autonomic, and endocrine functions in healthy volunteers. J Psychopharmacol 20: 756–770.

Sands SA and Morilak DA (1999) Expression of α1D adrenergic receptor messenger RNA in oxytocin- and corticotropin-releasing hormone-synthesizing neurons in the rat paraventricular nucleus. Neuroscience 91: 639–649.

Saper CB, Loewy AD, Swanson LW, et al. (1976) Direct hypothalamo-autonomic connections. Brain Res 117: 305–312.

Sara SJ and Devauges V (1988) Priming stimulation of locus coeruleus facilitates memory retrieval in the rat. Brain Res 438: 299–303.

Sasa M, Munekiyo H, Ikeda S, et al. (1974) Noradrenaline-mediated inhi-bition by locus coeruleus of spinal trigeminal neurons. Brain Res 80: 443–460.

Sasa M and Yoshimura N (1994) Locus coeruleus noradrenergic neurons as a micturition center. Microsc Res Tech 29: 226–230.

Sasaki M, Shibata E, Tohyama K, et al. (2006) Neuromelanin magnetic resonance imaging of locus coeruleus and substantia nigra in Parkin-son‘s disease. Neuroreport 17: 1215–1218.

Sawchenko PE and Swanson LW (1982) Immunohistochemical identifi-cation 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 205: 260–272.

Sawynok J and Reid A (1986) Role of ascending and descending nor-adrenergic pathways in the antinociceptive effect of clonidine and baclofen. Brain Res 386: 341–350.

Schambra UB, Mackensen GB, Stafford-Smith M, et al. (2005) Neu-ron specific α-adrenergic receptor expression in human cerebellum: Implications for emerging cerebellar roles in neurologic disease. Neu-roscience 135: 507–523.

Scheer FA, Ter Hors GJ, Van Der Vliet J, et al. (2001) Physiological and anatomic evidence for regulation of the heart by suprachiasmatic nucleus in rats. Am J Physiol Heart Circ Physiol 280: H1391–H1399.

Scheer FA, Van Doormen LJ and Buijs RM (1999) Light and diurnal cycle affect human heart rate: Possible role for the circadian pacemaker. J Biol Rhythms 14: 202–212.

Scheibner J, Trendelenburg AU, Hein L, et al. (2001) Alpha 2-adrenocep-tors modulating neuronal serotonin release: A study in alpha 2A-adre-noceptor-deficient mice. Br J Pharmacol 132: 925–933.

Scheibner J, Trendelenburg AU, Hein L, et al. (2002) Alpha 2-adrenocep-tors in the enteric nervous system: A study in alpha 2A-adrenoceptor-deficient mice. Br J Pharmacol 135: 697–704.

Scheinin M, Lomasney JW, Hayden-Hixson DM, et al. (1994) Distribu-tion of α2-adrenergic receptor subtype gene expression in rat brain. Mol Brain Res 21: 133–149.

Schofield SPM and Everitt BJ (1981) The organisation of catecholamine-containing neurons in the brain of the rhesus monkey (Macaca mulatta). J Anat 132: 391–418.

Schuerger RJ and Balaban CD (1993) Immunohistochemical demonstra-tion of regionally selective projections from locus coeruleus to the vestibular nuclei in rats. Exp Brain Res 92: 351–359.

Schwartz JRL and Roth T (2008) Neurophysiology of sleep and wakeful-ness: Basic science and clinical implications. Curr Neuropharmacol 6: 367–378.

Séguéla P, Watkins KC, Geffard M, et al. (1990) Noradrenaline axon ter-minals in adult rat neocortex: An immunocytochemical analysis in serial thin sections. Neuroscience 35: 249–264.

Semo M, Muñoz Llamosas M, Foster RG, et al. (2005) Melanopsin (Opn4) positive cells in the cat retina are randomly distributed across the ganglion cell layer. Vis Neurosci 22: 111–116.

Senba E, Tohyama M, Shiosaka H, et al. (1981) Experimental and mor-phological studies of the noradrenaline innervations in the nucleus tractus spinal nervi trigemini of the rat with special reference to their fine structures. Brain Res 206: 39–50.

Sesack SR, Deutch AY, Roth RH, et al. (1989) Topographical organiza-tion of the efferent projections of the medial prefrontal cortex in the

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 691

rat: An anterograde tract-tracing study with phaseolus vulgaris leuco-agglutinin. J Comp Neurol 290: 213–242.

Shao YP and Sutin J (1991) Noradrenergic facilitation of motor neurons: Localization of adrenergic receptors in neurons and nonneuronal cells in the trigeminal motor nucleus. Exp Neurol 114: 216–227.

Shefchyk SJ (2001) Sacral spinal interneurones and the control of urinary bladder and urethral striated sphincter muscle function. J Physiol 533: 57–63.

Shen H, Peri KG, Deng XF, et al. (2000) Distribution of alpha1-adreno-ceptor subtype proteins in different tissues of neonatal and adult rats. Can J Physiol Pharmacol 78: 237–243.

Shi TJ, Winzer-Serhan U, Leslie F, et al. (1999) Distribution of alpha2-adrenoceptor mRNAs in the rat lumbar spinal cord in normal and axotomized rats. Neuroreport 10: 2835–2839.

Shibata E, Sasaki M, Tohyama K, et al. (2006) Age-related changes in locus coeruleus on neuromelanin magnetic resonance imaging at 3 tesla. Magn Reson Med Sci 5: 197–200.

Shields AD, Wand Q and Winder DG (2009) Alpha2A-adrenergic recep-tors heterosynaptically regulate glutamatergic transmission in the bed nucleus of the stria terminalis. Neuroscience 163: 339–351.

Silberman Y, Ariwodola OJ and Weiner JL (2012) β1-adrenoceptor activation is required for ethanol enhancement of lateral paracapsular GABAergic synapses in the basolateral amygdala. J Pharmacol Exp Ther 343:451–459.

Simon H, Le Moal M, Stinus L, et al. (1979a) Efferents and afferents of the ventral tegmental-A10 region studied after local injection of [3H]leucine and horseradish peroxidase. Brain Res 178: 17–40.

Simon H, Le Moal M, Stinus L, et al. (1979b) Anatomical relationships between the ventral mesencephalic tegmentum-A10 region and the locus coeruleus as demonstrated by anterograde and retrograde trac-ing techniques. J Neural Transm 44: 77–86.

Singewald N and Philippu A (1998) Release of neurotransmitters in the locus coeruleus. Prog Neurobiol 56: 237–267.

Singewald N, Kaehler ST and Philippu A (1999) Noradrenaline release in the locus coeruleus of conscious rats triggered by drugs, stress and blood pressure changes. Neuroreport 10: 1583–1587.

Sirieix C, Gervasoni D, Luppi PH, et al. (2012) Role of the lateral paragigan-tocellular nucleus in the network of paradoxical (REM) sleep: An elec-trophysiological and anatomical study in the rat. PLoS One 7: e28724.

Sluka KA and Westlund KN (1992) Spinal projections of the locus coe-ruleus and the nucleus subcoeruleus in the Harlan and the Sasco Sprague-Dawley rat. Brain Res 579: 67–73.

Smith MS, Schambra UB, Wilson KH, et al. (1995) Alpha 2-adrener-gic 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 34: 109–117.

Smith MS, Schambra UB, Wilson KH, et al. (1999) α1-adrenergic recep-tors in human spinal cord: Specific localized expression of mRNA encoding α1-adrenergic receptor subtypes at four distinct levels. Mol Brain Res 63: 254–261.

Southwick SM, Krystal JH, Morgan CA, et al. (1993) Abnormal noradren-ergic function in posttraumatic stress disorder. Arch Gen Psychiatry 50: 266–274.

Spencer SE, Sawyer WB, Wada H, et al. (1990) CNS projections to the pterygopalatine parasympathetic preganglionic neurons in the rat: A retrograde transneuronal viral cell body labeling study. Brain Res 534: 149–169.

Starke K (2001) Presynaptic autoreceptors in the third decade: Focus on α2-adrenoceptors. J Neurochem 78: 685–693.

Steininger TL, Gong H, McGinty D, et al. (2001) Subregional organiza-tion of preoptic area/anterior hypothalamic projections to arousal-related monoaminergic cell groups. J Comp Neurol 429: 638–653.

Sterpenich V, D‘Argembeau A, Desseilles M, et al. (2006) The locus coe-ruleus is involved in the successful retrieval of emotional memories in humans. J Neurosci 26: 7416–7423.

Stock G, Rupprecht U, Stumpf H, et al. (1981) Cardiovascular changes during arousal elicited by stimulation of amygdala, hypothalamus and locus coeruleus. J Auton Nerv Syst 3: 503–510.

Stone A, Quartermain D, Lin Y, et al. (2006) Central α1-adrenergic sys-tem in behavioural activity and depression. Biochem Pharmacol 73: 1063–1075.

Stone EA and Ariano MA (1989) Are glial cells targets of the central norad-renergic system? A review of the evidence. Brain Res Rev 14: 297–309.

Sugiyama D, Hur SW, Pickering AE, et al. (2012) In vivo patch-clamp recording from locus coeruleus neurones in the rat brainstem. J Physiol 590: 2225–2231.

Sullivan RM, Wilson DA, Lemon C, et al. (1994) Bilateral 6-OHDA lesions of the locus coeruleus impair associative olfactory learning in newborn rats. Brain Res 643: 306–309.

Summers RJ, Papaionnaou M, Harris S, et al. (1995) Expression of beta 3-adrenoceptor mRNA in rat brain. Br J Pharmacol 116: 2547–2548.

Sun M-K (1995) Central neural organization and control of sympathetic nervous system in mammals. Prog Neurobiol 47: 157–233.

Swanson LW and Sawchenko PE (1980) Paraventricular nucleus: A site for the integration of neuroendocrine and autonomic mechanisms. Neuroendocrinology 31: 410–417.

Swanson LW and Sawchenko PE (1983) Hypothalamic integration: Orga-nization of the paraventricular and supraoptic nuclei. Ann Rev Neu-rosci 6: 269–324.

Szabadi E (1993) Functional neuroanatomy. In: RE Kendell and AK Zeally (eds) Companion to Psychiatric Studies, Fifth Edition. Edin-burgh: Churchill Livingstone, pp. 77–104.

Szabadi E (2006) Drugs for sleep disorders: Mechanisms and therapeutic prospects. Br J Clin Pharmacol 61: 761–766.

Szabadi E (2012) Modulation of physiological reflexes by pain: Role of the locus coeruleus. Front Integr Neurosci 6: 1–15.

Szabadi E and Bradshaw CM (1988) Biological markers for anxiety state. In: Granville-Grosman K (ed) Recent Advances in Clinical Psychia-try, vol. 6. Edinburgh: Churchill Livingstone, pp. 69–99.

Szabadi E and Bradshaw CM (1991) Adrenoceptors: Structure, Mecha-nisms, Function. Basel: Birkhäuser.

Szabadi E and Tavernor S (1999) Hypo- and hypersalivation induced by psychoactive drugs. Incidence, mechanisms and therapeutic implica-tions. CNS Drugs 11: 449–466.

Szabadi E, Biswas A, Langley R, et al. (2010) Effects of clonidine on responses of the human pupil to monochromatic light stimuli. Br J Clin Pharmacol 70: 300–301.

Szabadi E, Bradshaw CM and Nahorski SR (1985) Pharmacology of Adrenoceptors. Basingstoke: Macmillan.

Szot P, White SS, Greenup JL, et al. (2005) Alpha1-adrenoceptor in human hippocampus: Binding and receptor subtype mRNA expres-sion. Brain Res Mol Brain Res 139: 367–371.

Szymusiak R and McGinty D (2008) Hypothalamic regulation of sleep and arousal. Ann N Y Acad Sci 1129: 275–286.

Takakura ACT, Moreira TS, De Luca LA Jr, et al.(2003) Central α2 adren-ergic receptors and cholinergic-induced salivation in rats. Brain Res Bull 59: 383–386.

Takano T, Kubota Y, Wanaka A, et al. (1989) β-Adrenergic receptors in the vasopression-containing neurons in the paraventricular and supra-optic nucleus of the rat. Brain Res 499: 174–178.

Takauchi S, Yamauchi S, Morimura Y, et al. (1995) Coexistence of Pick bodies and atypical Lewy bodies in the locus coeruleus neurons of Pick‘s disease. Acta Neuropathol 90: 93–100.

Talley EM, Rosin DL, Lee A, et al. (1996) Distribution of alpha 2A-adren-ergic receptor-like immunoreactivity in the rat central nervous sys-tem. J Comp Neurol 372: 111–134.

Taneja P, Ogier M, Brooks-Haris G, et al. (2009) Pathophysiology of locus ceruleus neurons in a mouse model of Rett syndrome. J Neuro-sci 29: 12187–12195.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

692 Journal of Psychopharmacology 27(8)

Tavares A, Handy DE, Bogdanova NN, et al. (1996) Localization of α2A- and α2B-adrenergic receptor subtypes in brain. Hypertension 27: 449–455.

Ter Horst GJ, Toes GJ and Van Willigen JD (1991) Locus coeruleus pro-jections to the dorsal motor vagus nucleus in the rat. Neuroscience 45: 153–160.

Thapan K, Arendt J and Skene DJ (2001) An action spectrum for melato-nin suppression: Evidence for a novel non-rod, non-cone photorecep-tor system in humans. J Physiol 535: 261–267.

Thompson HS (1996) Autonomic control of the pupil. In: Robertson D, Low PA and Polinsky RJ (eds) Primer on the Autonomic Nervous System. San Diego: Academic Press, pp. 74–79.

Tillet Y and Kitahama K (1998) Distribution of central catecholaminergic neurons: A comparison between ungulates, humans and other species. Histol Histopathol 13: 1163–1177.

Tillet Y and Thibault J (1989) Catecholamine-containing neurons in the sheep brainstem and diencephalon: Immunohistochemical study with tyrosine hydroxylase (TH) and dopamine-beta-hydroxylase (DBH) antibodies. J Comp Neurol 290: 69–104.

Tomlinson BE, Irving D and Blessed G (1981) Cell loss in the locus coeruleus in senile dementia of Alzheimer‘s type. J Neurol Sci 49: 419–428.

Tóth IE, Boldogkői Z, Medveczky I, et al. (1999) Lacrimal preganglionic neurons form a subdivision of the superior salivatory nucleus of rat: Transneuronal labelling by pseudorabies virus. J Auton Nerv Syst 77: 45–54.

Trulson ME and Jacobs BL (1979) Raphe unit activity in freely moving cats: Correlation with level of behavioural arousal. Brain Res 163: 135–150.

Tsai JW, Hannibal J, Hagiware G, et al. (2009) Melanopsin as a sleep modulator: Circadian gating of the direct effects of light on sleep and altered homeostasis in Opn4–/– mice. PLoS Biol. 7: e1000125.

Tsuruoka M, Matsutani K, Maeda M, et al. (2003) Coeruleotrigeminal inhibition of nociceptive processing in the rat trigeminal subnucleus caudalis. Brain Res 993: 146–153.

Unnerstall JR, Kopajtic TA and Kuhar MJ (1984) Distribution of alpha 2 agonist binding sites in the rat and human central nervous system: Analysis of some functional, anatomic correlates of the pharmaco-logic effects of clonidine and related adrenergic agents. Brain Res 319: 69–101.

Ursino MG, Vasina V, Raschi E, et al. (2009) The β3-adrenoceptor as a therapeutic target: Current perspectives. Pharmacol Res 59: 221–234.

Uschakov A, Grivel J, Cvetkovic-Lopes V, et al. (2011) Sleep-deprivation regulates α-2 adrenergic responses of rat hypocretin/orexin neurons. PLoS ONE 6: e16672.

Valentino RJ and Van Bockstaele E (2008) Convergent regulation of locus coeruleus connectivity as an adaptive response to stress. Eur J Phar-macol 83: 194–203.

Van Bockstaele EJ (1998) Morphological substrates underlying opioid, epinephrine and γ-aminobutyric acid inhibitory actions in the rat locus coeruleus. Brain Res Bull 47: 1–15.

Van Bockstaele EJ and Aston-Jones G (1992) Collateralized projections from neurons in the rostral medulla to the nucleus locus coeruleus, the nucleus of the solitary tract and the periaqueductal gray. Neurosci-ence 49: 653–668.

Van Bockstaele EJ and Aston-Jones G (1995) Integration in the ventral medulla and coordination of sympathetic, pain and arousal functions. Clin Exp Hypertens 17: 153–165.

Van Bockstaele EJ, Bajic D, Proudfit H, et al. (2001) Topographic archi-tecture of stress-related pathways targeting the noradrenergic locus coeruleus. Physiol Behav 73: 273–283.

Van Bockstaele EJ, Colago EE and Aicher S (1998a) Light and electron microscopic evidence for topographic and monosynaptic projections from neurons in the ventral medulla to noradrenergic dendrites in the rat locus coeruleus. Brain Res 784: 123–138.

Van Bockstaele EJ, Colago EEO and Valentino RJ (1998b) Amygdaloid corticotrophin-releasing factor targets locus coeruleus dendrites: Sub-strate for the co-ordination of emotional and cognitive limbs of the stress response. J Neurosci 10: 743–757.

Van Bockstaele EJ, Pieribone VA and Aston-Jones G (1989) Diverse afferents converge on the nucleus paragigantocellularis in the rat ventrolateral medulla: Retrograde and anterograde tracing studies. J Comp Neurol 290: 561–584.

Van Dongen PA (1981) The central noradrenergic transmission and the locus coeruleus: A review of the data, and their implications for neu-rotransmission and neuromodulation. Prog Neurobiol 16: 117–143.

Van Esseveldt KE, Lehman MN and Boer GJ (2000) The suprachiasmatic nucleus and the circadian time-keeping system revisited. Brain Res Rev 33: 34–77.

Van Waarde, Visser TJ, Elsinga PH, et al. (1997) Imaging beta-adreno-ceptors in the human brain with (S )-1‘-[18F]Fluorocarazol. J Nucl Med 38: 934–939.

VanderMaelen CP and Aghajanian GK (1980) Intracellular studies show-ing modulation of facial motoneurone excitability by serotonin. Nature 287: 346–347.

Vandewalle G, Schmidt C, Albouy G, et al. (2007) Brain responses to violet, blue and green monochromatic light exposures in humans: Prominent role of blue light and the brainstem. PLoS ONE 11: e1247.

Vazey EM and Aston-Jones G (2012) The emerging role of norepineph-rine in cognitive dysfunctions in Parkinson‘s disease. Front Behav Neurosci 6: 1–6.

Verleye M and Bernet F (1983) Behavioural effects of lesions of the cen-tral noradrenergic bundle in the rat. Pharmacol Biochem Behav 19: 407–414.

Vertes RP and Kocsis B (1994) Projections of the dorsal raphe nucleus to the brainstem: PHA-L analysis in the rat. J Comp Neurol 340: 11–26.

Vetrivelan R, Chang C and Lu J (2011) Muscle tone regulation during REM sleep: Neural circuitry and clinical significance. Arch Ital Biol 149: 348–366.

Viemari JC, Roux JC, Tryba AK, et al. (2005) Mecp2 deficiency dis-rupts norepinephrine and respiratory systems in mice. J Neurosci 25: 11521–11530.

Vogt M (1954) The concentration of sympathin in different parts of the central nervous system under normal conditions and after administra-tion of drugs. J Physiol 123: 451–481.

Voisin DL, Guy N, Chalus M, et al. (2005) Nociceptive stimulation acti-vates locus coeruleus neurones projecting to the somatosensory thala-mus in the rat. J Physiol 566: 929–937.

Voituron N, Zanella S, Menuet C, et al. (2010) Early abnormalities of post-sigh breathing in a mouse model of Rett syndrome. Respir Physiol Neurobiol 170: 173–182.

Von Euler US (1951) The nature of adrenergic nerve mediators. Pharma-col Rev 3: 247–277.

Wallace DM, Magnuson DJ and Gray TS (1989) The amygdalo-brainstem pathway: Selective innervation of dopaminergic, noradrenergic and adrenergic cells in the rat. Neurosci Lett 97: 252–258.

Wanaka A, Kiyama H, Murakami T, et al. (1989) Immunocytochemical localization of β-adrenergic receptors in the rat brain. Brain Res 485: 125–140.

Wang LN, Zhu MW, Feng YQ, et al. (2006) Pick‘s disease with Pick bod-ies combined with progressive supranuclear palsy without tuft-shaped astrocytes: A clinical, neuroradiologic and pathological study of an autopsied case. Neuropathology 26: 222–230.

Warner JJ (2001) Atlas of Neuroanatomy. Boston: Butterworth-Heinemann.Watson M and McElligott JG (1984) Cerebellar noradrenaline depletion

and impaired acquisition of specific locomotor tasks in rats. Brain Res 296: 129–138.

West WL, Yeomans DC and Proudfit HK (1993) The function of norad-renergic neurons in mediating antinociception induced by electrical stimulation of the locus coeruleus in two different sources of Sprague-Dawley rats. Brain Res 626: 127–135.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Szabadi 693

Westlund KN and Coulter JD (1980) Descending projections of the locus coeruleus and subcoeruleus/medial parabrachial nuclei in monkey: Axonal transport studies and dopamine-beta-hydroxylase immunocy-tochemistry. Brain Res 2: 235–264.

Westlund KN, Bowker RM, Ziegler MG, et al. (1983) Noradrenergic pro-jections to the spinal cord of the rat. Brain Res 263: 15–31.

Westlund KN, Bowker RM, Ziegler MG, et al. (1984) Origins and termi-nations of descending noradrenergic projections to the spinal cord of the monkey. Brain Res 292: 1–16.

Westlund KN, Zhang D, Carlton SM, et al. (1991) Noradrenergic innervation of somatosensory thalamus and spinal cord. Prog Brain Res 88: 77–88.

White SR, Fung SJ and Barnes CD (1991) Norepinephrine effects on spi-nal motoneurons. Prog Brain Res 88: 343–350.

Wier WG, Zang WJ, Lamont C, et al. (2009) Sympathetic neurogenic Ca2+ signalling in rat arteries: ATP, noradrenaline and neuropeptide Y. Exp Physiol 94: 31–37.

Wilhelm B, Giedke H, Lϋdtke H, et al.(2001) Daytime variations in cen-tral nervous system activation measured by a pupillographic sleepi-ness test. J Sleep Res 10: 1–7.

Williams AM and Morilak DA (1997) Alpha1B adrenoceptors in rat para-ventricular nucleus overlap with, but do not mediate, the induction of c-Fos expression by osmotic or restraint stress. Neuroscience 76: 901–913.

Williams AM, Nguyen ML and Morilak DA (1997) Co-localization of alpha1D adrenergic receptor mRNA with mineralocorticoid and glu-cocorticoid receptor mRNA in rat hippocampus. J Neuroendocrinol 9: 113–119.

Willis WD and Westlund KN (1997) Neuroanatomy of the pain system and the pathways that modulate pain. J Clin Neurophysiol 14: 2–31.

Wise RA (2004) Dopamine, learning and motivation. Nat Rev Neurosci 5: 483–494.

Wu MF, Gulyani SA, Yau E, et al. (1999) Locus coeruleus neurons: Ces-sation of activity during cataplexy. Neuroscience 91: 1389–1399.

Xi M, Fung SJ, Zhang J, et al. (2012) The amygdala and the pedunculo-pontine tegmental nucleus: Interactions controlling active (rapid eye movement) sleep. Exp Neurol 238: 44–51.

Xu ZQ, Shi TJ and Hökfelt T (1998) Galanin/GMAP- and NPY-like immunoreactivities in locus coeruleus and noradrenergic nerve ter-minals in the hippocampal formation and cortex with notes on the galanin-R1 and -R2 receptors. J Comp Neurol 392: 227–251.

Yaïci E-D, Rampin O, Calas A, et al. (2002) α2a and α2c adrenoceptors on spinal neurons controlling penile erection. Neuroscience 114: 945–960.

Yamanaka A, Muraki Y, Tsujino N, et al. (2003) Regulation of orexin neurons by the monoaminergic and cholinergic systems. Biochem Biophys Res Commun 303: 120–129.

Yamazato M, Sakima A, Nakazato J, et al. (2001) Hypotensive and seda-tive effects of clonidine injected into the rostral ventrolateral medulla of conscious rats. Am J Physiol Regulatory Integrative Comp Physiol 281: 1868–1876.

Yang Y, Beyreuther K and Schmitt HP (1999) Spatial analysis of the neu-ronal density of aminergic brainstem nuclei in primary neurodegen-

erative and vascular dementia: A comparative immunocytochemical and quantitative study using a graph method. Anal Cell Pathol 19: 125–138.

Yokoo H, Kobayashi H, Minami S, et al. (2000) alpha(1)-Adrenergic receptor subtypes in rat cerebral microvessels. Brain Res 878: 183–187.

Yoshimura M, Higashi H and Nishi S (1985) Noradrenaline mediates slow excitatory synaptic potentials in rat dorsal raphe neurons in vitro. Neurosci Lett 61: 305–310.

Yoshimura M, Polosa C and Nishi S (1987) Slow EPSP and the depolar-izing action of noradrenaline on sympathetic preganglionic neurons. Brain Res 414: 138–142.

Yoshimura N, Sasa M, Yoshida O, et al. (1990a) Alpha 1-adrenergic receptor- mediated excitation from the locus coeruleus of the sacral parasympathetic preganglionic neuron. Life Sci 47: 789–797.

Yoshimura N, Sasa M, Yoshida O, et al. (1990b) Mediation of micturition reflex by central norepinephrine from the locus coeruleus in the cat. J Urol 143: 840–843.

Yoss RE, Moyer NJ and Hollenhurst RW (1970) Pupil size and spontane-ous pupillary waves associated with alertness, drowsiness, and sleep. Neurology 20: 545–554.

Young WS and Kuhar MJ (1980) Noradrenergic α1 and α2 receptors: Light microscopic autoradiographic localization. Proc Natl Acad Sci USA 77: 1696–1700.

Zagon A and Smith AD (1993) Monosynaptic projections from the rostral ventrolateral medulla oblongata to identified sympathetic pregangli-onic neurons. Neuroscience 54: 729–743.

Zarow C, Lyness SA, Mortimer JA, et al. (2003) Neuronal loss is greater in the locus coeruleus than nucleus basalis and substantia nigra in Alzheimer and Parkinson diseases. Arch Neurol 60: 337–341.

Zecharia AY, Yu X, Götz T, et al. (2012) GABAergic inhibition of his-taminergic neurons regulates active waking but not the sleep-wake switch or propofol-induced loss of consciousness. J Neurosci 32: 13062–13075.

Zeitzer JM, Dijk DJ, Kronauer R, et al. (2000) Sensitivity of the human circadian pacemaker to nocturnal light: Melatonin phase resetting and suppression. J Physiol 526: 695–702.

Zhang C, Guo Y-Q, Qiao J-T, et al. (1998a) Locus coeruleus modulates thalamic nociceptive responses via adrenoceptors. Brain Res 784: 116–122.

Zhang K-M, Wang X-M, Peterson AM, et al. (1998b) α2-adrenoceptors modulate NMDA-evoked responses of neurons in superfi-cial and deeper dorsal horn of the medulla. J Neurophysiol 80: 2210–2214.

Zhang X, Su J, Rojas A, et al. (2010) Pontine norepinephrine defects in Mecp2-null mice involve deficient expression of dopamine beta-hydroxylase but not a loss of catecholaminergic neurons. Biochem Biophys Res Commun 394: 285–290.

Zoghbi HY, Milstien S, Butler IJ, et al. (1989) Cerebrospinal fluid bio-genic amines and biopterin in Rett syndrome. Ann Neurol 25: 56–60.

Zweig RM, Cardillo JE, Cohen M, et al. (1993) The locus coeruleus and dementia in Parkinson‘s disease. Neurology 43: 986–991.

at Queens University on February 12, 2014jop.sagepub.comDownloaded from

Journal of Psychopharmacology27(10) 964

© The Author(s) 2013Reprints and permissions: sagepub.co.uk/journalsPermissions.navDOI: 10.1177/0269881113504545jop.sagepub.com

The publisher would like to apologize for the error that occurred in Issue 27/8 (August 2013), the Review article by Elemer Szabadi included Figures for the paper in black and white. They should have been published in colour as in the online version: http://jop.sage-pub.com/content/27/8/659.full

The Review article by contained Samuel R Chamberlain and Trevor W Robbins included Figure 6 for the paper in black and white. It should have been published in colour as in the online version: http://jop.sagepub.com/content/27/8/694.full

The article by contained A Pringle, C McCabe, PJ Cowen, and CJ Harmer should have been published as a Review. The article included Figure 1 for the paper in black and white. It should have been published in colour as in the online version: http://jop.sagepub.com/content/27/8/719.full

504545 JOP271010.1177/0269881113504545Journal of PsychopharmacologyErratum2013

Erratum