Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 Aug 15.
Published in final edited form as: Brain Res. 2016 Mar 9;1645:75–78. doi: 10.1016/j.brainres.2016.03.001

Locus Coeruleus: From Global Projection System to Adaptive Regulation of Behavior

Gary Aston-Jones 1,, Barry Waterhouse 2
PMCID: PMC4969192  NIHMSID: NIHMS767163  PMID: 26969408

Abstract

The brainstem nucleus locus coeruleus (LC) is a major source of norepinephrine (NE) projections throughout the CNS. This important property was masked in very early studies by the inability to visualize endogenous monoamines. The development of monoamine histofluorescence methods by Swedish scientists led to a plethora of studies, including a paper published in Brain Research by Loizou in 1969. That paper was highly cited (making it a focal point for the 50th anniversary issue of this journal), and helped to spark a large and continuing set of investigations to further refine our understating of the LC-NE system and its contribution to brain function and behavior. This paper very briefly reviews the ensuing advances in anatomical, physiological and behavioral aspects of the LC-NE system. Although its projections are ubiquitously present throughout the CNS, recent studies find surprising specificity within the organizational and operational domains of LC neurons. These and other findings lead us to expect that future work will unmask additional features of the LC-NE system and its roles in normative and pathological brain and behavioral processes.

Keywords: norepinephrine, noradrenaline, locus coeruleus, neural projections, arousal, attention

History of LC and Brain Norepinephrine

The locus coeruleus (LC) is one of the smallest but most extensively projecting nuclei in brain. It was first recognized in published material by Vicq-d’Azyr in 1786 (Tubbs et al., 2011). Russell notes in his extensive 1955 review (Russell, 1955) that the Wenzel bothers described LC in 1811 as a pigmented set of neurons in the dorsorostral tegmentum of human brain. Although unusual in its pigmentation, little else about LC garnered attention until the development of the Falck-Hillarp histofluoresence technique (Falck and Hillarp, 1959). This method allowed unprecedented visualization of tissue monoamines – suddenly, one could see cells, fibers and varicosities of the ubiquitous brain monoamines. Notable among these was the LC system. LC neurons in rat appeared to be nearly entirely aminergic, and a series of lesion and other studies revealed a small population of neurons that provided surprisingly widespread projections throughout the neuraxis. This novel efferent organization was unknown previously because existing methods for tracing fiber projections lacked ability to reveal the fine, unmyelinated fibers of LC neurons. Thus, these histofluorescence results were transformative and ushered in a new structural motif - a broad efferent network emanating from a small brainstem nucleus. This implied a global function for LC neurons as multiple areas with contrasting functions all received inputs from LC.

The 1969 Brain Research paper by Loizou (Loizou, 1969) was an important extension of reports by Swedish researchers on central aminergic systems. By studying the effects of lesions of monoamine fiber pathways (including LC projections) on fluorescence distal and proximal to the lesions (Andén et al., 1966; Dahlström and Fuxe, 1964); they inferred the sources and trajectories of monoamine projections including those from LC. Loizou (Loizou, 1969) extended these findings by making electrolytic lesions of the LC nucleus and examining histofluorescent fibers compared to what the Swedish groups had reported. He found that LC lesions greatly reduced aminergic fluorescent staining in several brain regions, including profound decreases in fibers and terminals in the dorsomotor nucleus of the trigeminal nerve and medullary tractus solitarius, as well as in forebrain regions including amygdala and hippocampus. Thus, his results provided important direct confirmation that LC projections were widespread and apparently diffuse. Often overlooked is his finding that such LC lesions decreased NE fibers in Edinger-Westphal nucleus, consistent with recent reports of relationships between LC neural activity and pupil diameter (Aston-Jones and Cohen, 2005; Gilzenrat et al., 2010; Joshi et al., 2016).

Modern LC Anatomy

Following this seminal work, there was considerable effort devoted to obtaining a more detailed characterization of LC and its circuit connections. This included anterograde and retrograde tract tracing and selective staining of NE neurons and fibers using antibodies directed against the NE synthetic enzyme, dopamine-beta-hydroxylase (DBH). This work confirmed the uniqueness of LC; a compact cluster of NE neurons whose projections distribute broadly throughout the neuraxis, from spinal cord to neocortex (Segal and Landis, 1974a; 1974b; Swanson and Hartman, 1975); reviewed in (Foote et al., 1983).

Neurochemical Composition of LC - Further investigation demonstrated that virtually all neurons within rodent and primate LC contained DBH(Grzanna and Molliver, 1980) and therefore NE as LC’s primary transmitter. However, later identification of multiple peptides co-localized within LC neurons added new dimensions to what initially appeared to be a conventional one-transmitter system. Vassopressin, somatostatin, neuropeptide Y, enkephalin, neurotensin, corticotropin-releasing factor and galanin are among the variety of putative peptide transmitters found in LC neurons (reviewed in (Aston-Jones, 2004; Olpe and Steinmann, 1991)).

Efferent Topography of LC - Multiple reports beginning in the late 1970’s provided evidence of an efferent topography within LC, i.e., a spatial organization of cells in the LC nucleus with respect to terminal field targets (reviewed in – Berridge and Waterhouse, 2003). For example, cortically-projecting LC neurons are more prominent within the caudal portion of LC and these neurons project in a predominantly ipsilateral (>95%) manner (Waterhouse et al., 1983). By contrast cells projecting to sub-cortical structures exhibit a more pronounced bilateral distribution. Although early retrograde tracing and antidromic activation studies revealed that single LC neurons branch to innervate different brain regions (reviewed in (Aston-Jones, 2004)), other studies (Simpson et al., 1997; Steindler, 1981) indicate that individual LC neurons collateralize to innervate functionally related circuits. For example, LC neurons that project to trigeminal somatosensory cortex are more likely to co-innervate trigeminal somatosensory thalamus than non-somatosensory thalamic regions (Simpson et al., 1997). These findings indicate that LC neurons collateralize according to functional properties of targets.

More recent studies demonstrated segregation among LC neurons that project to primary motor cortex and sub-regions of the prefrontal cortex: anterior cingulate, orbitofrontal, and medial prefrontal (Chandler et al., 2014; 2013). Subsequent examination of the membrane properties and molecular phenotypes of LC-prefrontal and LC-motor cortical projection cells revealed differences in excitability and protein expression between these two groups. Combined, these data indicate that subsets of LC neurons may be capable of asynchronous release of NE in sub-regions of the cortex and distinct roles in executive function versus motor control.

Terminal Field Distribution of NE-containing Fibers - Although early work emphasized the diffuse nature of LC projections throughout the brain, subsequent investigations in monkey and human found substantial regional specificity of noradrenergic fiber distribution across cortical and subcortical structures. Within the primate visual system, NE fibers are more heavily represented in tecto-pulvinar-juxtastriate structures as contrasted with geniculo-striate or inferotemporate structures (Foote and Morrison, 1987) On the basis of this distinction Morrison and Foote (Morrison and Foote, 1986).suggested that, within the visual system, NE-LC fibers preferentially innervate regions involved in spatial analysis and visuomotor responses rather than areas involved in feature extraction or pattern analysis. Thus, on the basis of fiber distribution, the LC-NE system may potentially have a selective influence on specific dimensions of visual signal processing. Additional selectivity was demonstrated through analysis of DBH immunostained axonal varicosities across rat cortical sub-regions (Agster et al., 2013). These studies showed that the medial prefrontal region exhibits the highest density of NE varicosities of any cortical region examined. These results provide further anatomical evidence that LC-mediated release of NE is not uniform across the cortical mantle.

Synaptic vs Non-synaptic Release of NE – Volume Transmission - The anatomical relationships of varicosities on NE axons – the presumed transmitter release points - have been controversial (Descarries and Lapierre, 1973; Lapierre et al. 1973; Olschowka et al, 1981; Watkins et al, 1977). Early studies indicated that a high percentage of such varicosities were non-synaptic, leading to a volume transmission theory of NE signaling in the brain (Agnati et al, 1995). However, later work found a greater incidence of conventional synaptic contacts between NE terminals and targets, arguing for more a more specific cell to cell release and action of NE than implied by earlier studies (Olschowka et al., 1981; Papadopoulos and Parnavelas, 1990; Papadopoulos et al., 1987; 1989). It seems possible that both synaptic and non-synaptic modes of NE release occur.

Afferent Regulation of LC - Initial reports indicated that the LC receives inputs from a broad array of CNS structures (Cedarbaum and Aghajanian, 1978). However, subsequent work found that the LC nucleus receives a restricted set of afferents, arising primarily from the ventrolateral and dorsomedial rostral medulla as well as hypothalamus (Aston-Jones et al., 1986). Subsequent studies documented a dense plexus of LC neuronal dendrites that extend far beyond the borders of the nucleus proper into pericoerulear zones (Shipley et al., 1996). Importantly, these pericoerulear regions are targeted by inputs from a variety of sources including prefrontal cortex, amygdala, lateral hypothalamus and dorsal raphe. Recent work using viral-genetic tracing methods confirmed that NE neurons in LC received inputs from a wide array of brain areas, and also exhibited a high divergence in output projections (Schwarz et al., 2015). Jodo and Aston-Jones (Jodo and Aston-Jones, 1997) found that PFC activity activates LC neurons presumably through pericoerulear dendrites, an important connection insofar as it links circuits involved in higher cognitive processes (e.g. working memory) with the LC-efferent path.

The pericoerulear zone also contains a dense collection of GABA neurons co-mingled with peri-LC dendrites (Aston-Jones et al., 2004). This region receives inputs from many brain regions, and resembles a pool of inhibitory interneurons that provide either feed-forward or feedback inhibition of LC-NE neurons. Increasing work is revealing additional transmitter-defined inputs to LC, including hypocretin/orexin, CRF, glutamate, and others (for review, see (Aston-Jones, 2004).

In addition, the transsynaptic retrograde tracer pseudorabies virus (PRV) has revealed circuit-level inputs to LC with functional implications. Studies using this tracer have connected the suprachiasmatic nucleus with the LC via relays in the hypothalamic dorsomedial nucleus including hypocretin neurons (Aston-Jones et al., 2001; Gompf and Aston-Jones, 2008). This was the first circuit identified in the brain for circadian regulation of arousal.

Effects of NE on target neurons

In early studies, application of NE by microiontophoresis, or stimulation of the LC, suppressed spontaneous neural activity in cerebellum (Bloom et al., 1971; Hoffer et al., 1971; 1973; Siggins et al., 1971a; 1971b), cerebral cortex (Armstrong-James and Fox, 1983; Stone, 1973), and elsewhere in brain (Segal and Bloom, 1974a; 1974b; Siggins and Gruol, 1986). These studies, many of which were published in Brain Research, were compelling in so far as they provided anatomical, biochemical, electrophysiological, and pharmacological evidence establishing NE as a putative inhibitory neurotransmitter at central synapses. Moreover, these investigations provided the foundation for more than 40 years of ensuing work leading to modern theories of the role of the LC-NE system in brain function and behavior.

Modulatory Effects of NE on Cells and Circuits - In the late 1970’s pioneering studies in monkey auditory cortex (Foote et al., 1975), hippocampus (Segal and Bloom, 1976) and cerebellum (Freedman et al., 1977) demonstrated a modulatory effect of NE, such that the spontaneous firing rates of cells were suppressed to a greater extent than stimulus-evoked discharge, thus yielding a net increase in “signal to noise” ratio. At levels of NE which had minimal or no effect on baseline firing, Woodward and colleagues found that stimulus-evoked excitation and inhibition in cerebellum and somatosensory cortex were increased above control levels during iontophoretic NE or LC stimulation (reviewed in (Berridge and Waterhouse, 2003)). These findings supported the idea that a prominent physiological function of central NE might be to enhance the efficacy of both excitatory and inhibitory synaptic transmission rather than directly suppress cell firing (Woodward et al., 1979). Additional work showed that such actions can lead to selective alteration of the feature extraction properties of individual sensory neurons (Ciombor et al., 1999; Doucette et al., 2007; Kasamatsu and Heggelund, 1982; McLean and Waterhouse, 1994). Subsequent multi-neuron recordings showed that LC output can simultaneously modulate neurons at multiple sites along sensory pathways (Devilbiss et al., 2006; Devilbiss and Waterhouse, 2011).

Activity of LC Neurons and LC Function

Early recordings found that LC neurons were most active during waking and decreased activity in sleep (Aston-Jones and Bloom, 1981a; Hobson et al., 1975). Subsequent work showed that LC neurons also phasically respond to salient stimuli that produce behavioral responses (Aston-Jones and Bloom, 1981b; Foote et al., 1980). More recent studies found that the tonic and phasic patterns of LC activity interact, such that LC neurons fire tonically at a moderate rate but are phasically activated by task-relevant cues during periods of focused attention. In contrast, high tonic LC activity without phasic responses occurs during periods of low utility when task attentiveness wanes. These and other results led to the Adaptive Gain Theory (Aston-Jones and Cohen, 2005), which proposes that LC neurons are phasically activated in response to decision outcome, helping to execute adaptive behavioral responses. In contrast, high tonic LC discharge in this model serves to disrupt low utility behavior and increase behavioral flexibility so that more adaptive strategies can be pursued.

Synopsis.

Overall, much has been learned since Loizou’s initial observations about the extent of projections from LC-NE neurons. Subsequent studies have shown that multiple inputs differentially engage the LC nuclear core and dendritic shell so as to increase NE transmission across broad expanses of CNS tissue in a selective manner. Experimental evidence and theoretical constructs argue that such NE release serves to facilitate neural network functions and optimize behavioral outcomes across a range of contingencies that require focused versus flexible attention (Aston-Jones and Cohen, 2005; Bouret and Sara, 2005).

Acknowledgments

This work was supported by PHS grants R01-MH092868, R01-MH101178 and R01-DA006214 and R01-DA017960.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Contributor Information

Gary Aston-Jones, Email: aston.jones@rutgers.edu.

Barry Waterhouse, Email: Waterhouse@DrexelMed.edu.

References

  1. Agster KL, Mejías-Aponte CA, Clark BD, Waterhouse BD. Evidence for a regional specificity in the density and distribution of noradrenergic varicosities in rat cortex. J Comp Neurol. 2013;521:2195–2207. doi: 10.1002/cne.23270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Andén NE, Dahlström A, Fuxe K, Larsson K, Olson L, Ungerstedt U. Ascending monoamine neurons to the telencephalon and diencephalon. Acta Physiol Scand. 1966;67:313–326. [Google Scholar]
  3. Armstrong-James M, Fox K. Effects of ionophoresed noradrenaline on the spontaneous activity of neurones in rat primary somatosensory cortex. J Physiol (Lond) 1983;335:427–447. doi: 10.1113/jphysiol.1983.sp014542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Aston-Jones G. Locus coeruleus, A5 and A7 noradrenergic cell groups. In: Paxinos G, editor. The Rat Nervous System. Elsevier Academic Press; San Diego: 2004. pp. 259–294. [Google Scholar]
  5. Aston-Jones G, Bloom FE. Activity of norepinephrine-containing locus coeruleus neurons in behaving rats anticipates fluctuations in the sleep-waking cycle. J Neurosci. 1981a;1:876–886. doi: 10.1523/JNEUROSCI.01-08-00876.1981. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Aston-Jones G, Bloom FE. Norepinephrine-containing locus coeruleus neurons in behaving rats exhibit pronounced responses to non-noxious environmental stimuli. J Neurosci. 1981b;1:887–900. doi: 10.1523/JNEUROSCI.01-08-00887.1981. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Aston-Jones G, Chen S, Zhu Y, Oshinsky ML. A neural circuit for circadian regulation of arousal. Nat Neurosci. 2001;4:732–738. doi: 10.1038/89522. [DOI] [PubMed] [Google Scholar]
  8. Aston-Jones G, Cohen JD. An integrative theory of locus coeruleus-norepinephrine function: adaptive gain and optimal performance. Annu Rev Neurosci. 2005;28:403–450. doi: 10.1146/annurev.neuro.28.061604.135709. [DOI] [PubMed] [Google Scholar]
  9. Aston-Jones G, Ennis M, Pieribone VA, Nickell WT, Shipley MT. The brain nucleus locus coeruleus: restricted afferent control of a broad efferent network. Science. 1986;234:734–737. doi: 10.1126/science.3775363. [DOI] [PubMed] [Google Scholar]
  10. Aston-Jones G, Zhu Y, Card JP. Numerous GABAergic afferents to locus ceruleus in the pericerulear dendritic zone: possible interneuronal pool. Journal of Neuroscience. 2004;24:2313–2321. doi: 10.1523/JNEUROSCI.5339-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Berridge CW, Waterhouse BD. The locus coeruleus-noradrenergic system: modulation of behavioral state and state-dependent cognitive processes. Brain Res Brain Res Rev. 2003;42:33–84. doi: 10.1016/s0165-0173(03)00143-7. [DOI] [PubMed] [Google Scholar]
  12. Bloom FE, Hoffer BJ, Siggins GR. Studies on norepinephrine-containing afferents to Purkinje cells of art cerebellum. I Localization of the fibers and their synapses. Brain Research. 1971;25:501–521. doi: 10.1016/0006-8993(71)90457-4. [DOI] [PubMed] [Google Scholar]
  13. Bouret S, Sara S. Network reset: a simplified overarching theory of locus coeruleus noradrenaline function. Trends in Neurosciences. 2005;28:574–582. doi: 10.1016/j.tins.2005.09.002. [DOI] [PubMed] [Google Scholar]
  14. Cedarbaum JM, Aghajanian GK. Afferent projections to the rat locus coeruleus as determined by a retrograde tracing technique. J Comp Neurol. 1978;178:1–16. doi: 10.1002/cne.901780102. [DOI] [PubMed] [Google Scholar]
  15. Chandler DJ, Gao WJ, Waterhouse BD. Heterogeneous organization of the locus coeruleus projections to prefrontal and motor cortices. Proceedings of the National Academy of Sciences. 2014;111:6816–6821. doi: 10.1073/pnas.1320827111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Chandler DJ, Lamperski CS, Waterhouse BD. Identification and distribution of projections from monoaminergic and cholinergic nuclei to functionally differentiated subregions of prefrontal cortex. Brain Research. 2013;1522:38–58. doi: 10.1016/j.brainres.2013.04.057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Ciombor KJ, Ennis M, Shipley MT. Norepinephrine increases rat mitral cell excitatory responses to weak olfactory nerve input via alpha-1 receptors in vitro. Neuroscience. 1999;90:595–606. doi: 10.1016/s0306-4522(98)00437-0. [DOI] [PubMed] [Google Scholar]
  18. Dahlström A, Fuxe K. Evidence for the existence of monoamine-containing neurons in the central nervous system. I demonstration of monoamines in the cell bodies of brain stem neurons. Acta Physiol Scand. 1964;62:5–55. [PubMed] [Google Scholar]
  19. Devilbiss DM, Page ME, Waterhouse BD. Locus ceruleus regulates sensory encoding by neurons and networks in waking animals. J Neurosci. 2006;26:9860–9872. doi: 10.1523/JNEUROSCI.1776-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Devilbiss DM, Waterhouse BD. Phasic and tonic patterns of locus coeruleus output differentially modulate sensory network function in the awake rat. J Neurophysiol. 2011;105:69–87. doi: 10.1152/jn.00445.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Doucette W, Milder J, Restrepo D. Adrenergic modulation of olfactory bulb circuitry affects odor discrimination. Learn Mem. 2007;14:539–547. doi: 10.1101/lm.606407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Falck B, Hillarp NA. On the cellular localization of catechol amines in the brain. Acta Anat (Basel) 1959;38:277–279. doi: 10.1159/000141530. [DOI] [PubMed] [Google Scholar]
  23. Foote SL, Aston-Jones G, Bloom FE. Impulse activity of locus coeruleus neurons in awake rats and monkeys is a function of sensory stimulation and arousal. Proc Natl Acad Sci USA. 1980;77:3033–3037. doi: 10.1073/pnas.77.5.3033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Foote SL, Bloom FE, Aston-Jones G. Nucleus locus ceruleus: new evidence of anatomical and physiological specificity. Physiol Rev. 1983;63:844–914. doi: 10.1152/physrev.1983.63.3.844. [DOI] [PubMed] [Google Scholar]
  25. Foote SL, Freedman R, Oliver AP. Effects of putative neurotransmitters on neuronal activity in monkey auditory cortex. Brain Research. 1975;86:229–242. doi: 10.1016/0006-8993(75)90699-x. [DOI] [PubMed] [Google Scholar]
  26. Foote SL, Morrison JH. Extrathalamic modulation of cortical function. Annu Rev Neurosci. 1987;10:67–95. doi: 10.1146/annurev.ne.10.030187.000435. [DOI] [PubMed] [Google Scholar]
  27. Freedman R, Hoffer BJ, Woodward DJ, Puro D. Interaction of norepinephrine with cerebellar activity evoked by mossy and climbing fibers. Exp Neurol. 1977;55:269–288. doi: 10.1016/0014-4886(77)90175-3. [DOI] [PubMed] [Google Scholar]
  28. Gilzenrat MS, Nieuwenhuis S, Jepma M, Cohen JD. Pupil diameter tracks changes in control state predicted by the adaptive gain theory of locus coeruleus function. Cognitive, affective & behavioral neuroscience. 2010;10:252–269. doi: 10.3758/CABN.10.2.252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Gompf HS, Aston-Jones G. Role of orexin input in the diurnal rhythm of locus coeruleus impulse activity. Brain Research. 2008;1224:43–52. doi: 10.1016/j.brainres.2008.05.060. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Grzanna R, Molliver ME. The locus coeruleus in the rat: an immunohistochemical delineation. Neuroscience. 1980;5:21–40. doi: 10.1016/0306-4522(80)90068-8. [DOI] [PubMed] [Google Scholar]
  31. Hobson JA, McCarley RW, Wyzinski PW. Sleep cycle oscillation: Reciprocal discharge by two brainstem neuronal groups. Science. 1975;189:55–58. doi: 10.1126/science.1094539. [DOI] [PubMed] [Google Scholar]
  32. Hoffer BJ, Siggins GR, Bloom FE. Studies on norepinephrine-containing afferents to Purkinje cells of rat cerebellum. II Sensitivity of Purkinje cells to norepinephrine and related substances administered by microiontophoresis. Brain Research. 1971;25:523–534. doi: 10.1016/0006-8993(71)90458-6. [DOI] [PubMed] [Google Scholar]
  33. Hoffer BJ, Siggins GR, Oliver AP, Bloom FE. Activation of the pathway from locus coeruleus to rat cerebellar Purkinje neurons: pharmacological evidence of noradrenergic central inhibition. J Pharmacol Exp Ther. 1973;184:553–569. [PubMed] [Google Scholar]
  34. Jodo E, Aston-Jones G. Activation of locus coeruleus by prefrontal cortex is mediated by excitatory amino acid inputs. Brain Research. 1997;768:327–332. doi: 10.1016/s0006-8993(97)00703-8. [DOI] [PubMed] [Google Scholar]
  35. Joshi S, Li Y, Kalwani RM, Gold JI. Relationships between Pupil Diameter and Neuronal Activity in the Locus Coeruleus, Colliculi, and Cingulate Cortex. Neuron. 2016;89:221–234. doi: 10.1016/j.neuron.2015.11.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Kasamatsu T, Heggelund P. Single cell responses in cat visual cortex to visual stimulation during iontophoresis of noradrenaline. Exp Brain Res. 1982;45:317–327. doi: 10.1007/BF01208591. [DOI] [PubMed] [Google Scholar]
  37. Loizou LA. Projections of the nucleus locus coeruleus in the albino rat. Brain Research. 1969;15:563–566. doi: 10.1016/0006-8993(69)90185-1. [DOI] [PubMed] [Google Scholar]
  38. McLean J, Waterhouse BD. Noradrenergic modulation of cat area 17 neuronal responses to moving visual stimuli. Brain Research. 1994;667:83–97. doi: 10.1016/0006-8993(94)91716-7. [DOI] [PubMed] [Google Scholar]
  39. Morrison JH, Foote SL. Noradrenergic and serotoninergic innervation of cortical, thalamic, and tectal visual structures in Old and New World monkeys. J Comp Neurol. 1986;243:117–138. doi: 10.1002/cne.902430110. [DOI] [PubMed] [Google Scholar]
  40. Olpe HR, Steinmann M. Responses of locus coeruleus neurons to neuropeptides. Prog Brain Res. 1991;88:241–248. doi: 10.1016/s0079-6123(08)63813-3. [DOI] [PubMed] [Google Scholar]
  41. Olschowka JA, Molliver ME, Grzanna R, Rice FL, Coyle JT. Ultrastructural demonstration of noradrenergic synapses in the rat central nervous system by dopamine-beta-hydroxylase immunocytochemistry. J Histochem Cytochem. 1981;29:271–280. doi: 10.1177/29.2.7019303. [DOI] [PubMed] [Google Scholar]
  42. Papadopoulos GC, Parnavelas JG. Distribution and synaptic organization of serotoninergic and noradrenergic axons in the lateral geniculate nucleus of the rat. J Comp Neurol. 1990;294:345–355. doi: 10.1002/cne.902940304. [DOI] [PubMed] [Google Scholar]
  43. Papadopoulos GC, Parnavelas JG, Buijs R. Monoaminergic fibers form conventional synapses in the cerebral cortex. Neuroscience Letters. 1987;76:275–279. doi: 10.1016/0304-3940(87)90414-9. [DOI] [PubMed] [Google Scholar]
  44. Papadopoulos GC, Parnavelas JG, Buijs RM. Light and electron microscopic immunocytochemical analysis of the noradrenaline innervation of the rat visual cortex. J Neurocytol. 1989;18:1–10. doi: 10.1007/BF01188418. [DOI] [PubMed] [Google Scholar]
  45. Russell GV. The nucleus locus coeruleus (dorsolateralis tegmenti) Tex Rep Biol Med. 1955;13:939–988. [PubMed] [Google Scholar]
  46. Schwarz LA, Miyamichi K, Gao XJ, Beier KT, Weissbourd B, DeLoach KE, Ren J, Ibanes S, Malenka RC, Kremer EJ, Luo L. Viral-genetic tracing of the input-output organization of a central noradrenaline circuit. Nature. 2015;524:88–92. doi: 10.1038/nature14600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Segal M, Bloom FE. The action of norepinephrine in the rat hippocampus. IV The effects of locus coeruleus stimulation on evoked hippocampal unit activity. Brain Research. 1976;107:513–525. doi: 10.1016/0006-8993(76)90141-4. [DOI] [PubMed] [Google Scholar]
  48. Segal M, Bloom FE. The action of norepinephrine in the rat hippocampus. I Iontophoretic studies. Brain Research. 1974a;72:79–97. doi: 10.1016/0006-8993(74)90652-0. [DOI] [PubMed] [Google Scholar]
  49. Segal M, Bloom FE. The action of norepinephrine in the rat hippocampus. II Activation of the input pathway. Brain Research. 1974b;72:99–114. doi: 10.1016/0006-8993(74)90653-2. [DOI] [PubMed] [Google Scholar]
  50. Segal M, Landis S. Afferents to the hippocampus of the rat studied with the method of retrograde transport of horseradish peroxidase. Brain Research. 1974a;78:1–15. doi: 10.1016/0006-8993(74)90349-7. [DOI] [PubMed] [Google Scholar]
  51. Segal M, Landis SC. Afferents to the septal area of the rat studied with the method of retrograde axonal transport of horseradish peroxidase. Brain Research. 1974b;82:263–268. doi: 10.1016/0006-8993(74)90603-9. [DOI] [PubMed] [Google Scholar]
  52. Shipley MT, Fu L, Ennis M, Liu WL, Aston-Jones G. Dendrites of locus coeruleus neurons extend preferentially into two pericoerulear zones. J Comp Neurol. 1996;365:56–68. doi: 10.1002/(SICI)1096-9861(19960129)365:1<56::AID-CNE5>3.0.CO;2-I. [DOI] [PubMed] [Google Scholar]
  53. Siggins GR, Gruol DL. Mechanisms of transmitter action in the vertabrate central nervous system. In: Bloom FE, editor. Handbook of Physiology. American Physiol. Soc; Bethesda, MD: 1986. pp. 1–114. [Google Scholar]
  54. Siggins GR, Hoffer BJ, Bloom FE. Studies on norepinephrine-containing afferents to Purkinje cells of rat cerebellum. 3. Evidence for mediation of norepinephrine effects by cyclic 3“,5-”adenosine monophosphate. Brain Research. 1971a;25:535–553. doi: 10.1016/0006-8993(71)90459-8. [DOI] [PubMed] [Google Scholar]
  55. Siggins GR, Oliver AP, Hoffer BJ, Bloom FE. Cyclic adenosine monophosphate and norepinephrine: effects on transmembrane properties of cerebellar Purkinje cells. Science. 1971b;171:192–194. doi: 10.1126/science.171.3967.192. [DOI] [PubMed] [Google Scholar]
  56. Simpson KL, Altman DW, Wang L, Kirifides ML, Lin RC, Waterhouse BD. Lateralization and functional organization of the locus coeruleus projection to the trigeminal somatosensory pathway in rat. J Comp Neurol. 1997;385:135–147. [PubMed] [Google Scholar]
  57. Steindler DA. Locus coeruleus neurons have axons that branch to the forebrain and cerebellum. Brain Research. 1981;223:367–373. doi: 10.1016/0006-8993(81)91149-5. [DOI] [PubMed] [Google Scholar]
  58. Stone TW. Pharmacology of pyramidal tract cells in the cerebral cortex. Noradrenaline and related substances Naunyn Schmiedebergs. Arch Pharmacol. 1973;278:333–346. doi: 10.1007/BF00501478. [DOI] [PubMed] [Google Scholar]
  59. Swanson LW, Hartman BK. The central adrenergic system. An immunofluorescence study of the location of cell bodies and their efferent connections in the rat utilizing dopamine-beta-hydroxylase as a marker. J Comp Neurol. 1975;163:467–505. doi: 10.1002/cne.901630406. [DOI] [PubMed] [Google Scholar]
  60. Tubbs RS, Loukas M, Shoja MM, Mortazavi MM, Cohen-Gadol AA. Félix Vicq d’Azyr (1746–1794): early founder of neuroanatomy and royal French physician., Child’s nervous system. ChNS : official journal of the International Society for Pediatric Neurosurgery. 2011 doi: 10.1007/s00381-011-1424-y. [DOI] [PubMed] [Google Scholar]
  61. Waterhouse BD, Lin CS, Burne RA, Woodward DJ. The distribution of neocortical projection neurons in the locus coeruleus. J Comp Neurol. 1983;217:418–431. doi: 10.1002/cne.902170406. [DOI] [PubMed] [Google Scholar]
  62. Woodward DJ, Moises HC, Waterhouse BD, Hoffer BJ, Freedman R. Modulatory actions of norepinephrine in the central nervous system. Fed Proc. 1979;38:2109–2116. [PubMed] [Google Scholar]

RESOURCES