Skip to main content Accessibility help
×
Hostname: page-component-7c8c6479df-xxrs7 Total loading time: 0 Render date: 2024-03-17T17:31:58.467Z Has data issue: false hasContentIssue false

Chapter 10 - Brainstem

Published online by Cambridge University Press:  22 February 2018

David L. Clark
Affiliation:
Ohio State University
Nash N. Boutros
Affiliation:
University of Missouri, Kansas City
Mario F. Mendez
Affiliation:
University of California, Los Angeles
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
The Brain and Behavior
An Introduction to Behavioral Neuroanatomy
, pp. 151 - 163
Publisher: Cambridge University Press
Print publication year: 2018

Access options

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

References

Select Bibliography

Arango, V., Underwood, M. D., and Mann, J. J. (1997). Biological alterations in the brainstem of suicides. In Mann, J. J. (Ed.), Suicide. Psychiatric Clinics of North America (vol. 20, pp. 581594). Philadelphia: Saunders.Google Scholar
Carpenter, M. B., and Sutin, J. (1983). Human Neuroanatomy. Baltimore, MD: Williams and Wilkins.Google Scholar
Lenz, F. A., Casey, K. L., Jones, E. G., and Willis, W. D. (2010). The Human Pain System. Cambridge, UK: Cambridge University Press.Google Scholar
Hales, R. E., and Ydofsky, S. C. (1987). Textbook of Neuropsychiatry. Washington DC: American Psychiatric Association Press.Google Scholar
Holstege, G., Bandler, R., and Saper, C. B. (1996). The Emotional Motor System. Progress in Brain Research (vol. 107). New York, NY: Elsevier.Google Scholar

References

Adell, A. (2015). Revisiting the role of raphe and serotonin in neuropsychiatric disorders. J. Gen. Physiol., 145(4), 257259. doi:10.1085/jgp.201511389Google Scholar
Aston-Jones, G., and Cohen, J. D. (2005). An integrative theory of locus coeruleus-norepinephrine function: Adaptive gain and optimal performance. Annu. Rev. Neurosci., 28, 403450. doi:10.1146/annurev.neuro.28.061604.135709Google Scholar
Aston-Jones, G., and Gold, J. I. (2009). How we say no: Norepinephrine, inferior frontal gyrus, and response inhibition. Biol. Psychiatry, 65, 548549. doi:10.1016/j.biopsych.2009.01.022CrossRefGoogle ScholarPubMed
Aston-Jones, G., Rajkowski, J., and Cohen, J. (1999). Role of locus coeruleus in attention and behavioral flexibility. Biol. Psychiatry, 46(9), 13091320. Retrieved from: www.ncbi.nlm.nih.gov/pubmed/10560036Google Scholar
Aston-Jones, G., Rajkowski, J., Lu, W., Zhu, Y., Cohen, J. D., and Morecraft, R. J. (2002). Prominent projections from the orbital prefrontal cortex to the locus ceruleus in monkey. Soc. Neurosci. Abstr., 28, 8689. doi:10.3389/fncom.2012.0000Google Scholar
Aston-Jones, G., Smith, R. J., Sartor, G. C., Moorman, D. E., Massi, L., Tahsili-Fahadan, P., and Richardson, K. A. (2010). Lateral hypothalamic orexin/hypocretin neurons: A role in reward-seeking and addiction. Brain Res., 1314, 7490. doi:10.1016/j.brainres.2009.09.106Google Scholar
Baker, K. G., Halliday, G. M., Halasz, P., Hornung, J. P., Geffen, L. B., Cotton, R. G., and Törk, I. (1991). Cytoarchitecture of serotonin-synthesizing neurons in the pontine tegmentum of the human brain. Synapse, 7(4), 301320. Retrieved from www.ncbi.nlm.nih.gov/pubmed/2042112Google Scholar
Bangasser, D. A., Wiersielis, K. R., and Khantsis, S. (2016). Sex differences in the locus coeruleus-norepinephrine system and its regulation by stress. Brain Res., 1641(B), 177188. doi:10.1016/j.brainres.2015.11.021Google Scholar
Benarroch, E. E. (2012). Periaqueductal gray: An interface for behavioral control. Pedunculopontine nucleus. Neurology, 78, 210217. doi:10.1212/WNL.0b0131823fcdeeGoogle Scholar
Bensaid, M., Michel, P. P., Clark, S. D., Hirsch, E. C., and François, C. (2016). Role of pedunculopontine cholinergic neurons in the vulnerability of nigral dopaminergic neurons in Parkinson’s disease. Exp. Neurol., 275(1), 209219. doi:10.1016/j.expneurol.2015.11.004CrossRefGoogle ScholarPubMed
Berridge, C. W., Devilbiss, D. M., Andrzejewski, M. E., Arnsten, A. F., Kelley, A. E., Schmeichel, B., … Spencer, R. C. (2006). Methylphenidate preferentially increases catecholamine neurotransmission within the prefrontal cortex at low doses that enhance cognitive function. Biol. Psychiatry, 60(10), 11111120. doi:10.1016/j.biopsych.2006.04.022CrossRefGoogle ScholarPubMed
Black, S. W., Yamanaka, A., and Kilduff, T. S. (2015). Challenges in the development of therapeutics for narcolepsy. Prog. Neurobiol., 152, 89113. doi:10.1016/j.pneurobio.2015.12.002Google Scholar
Boyer, E. W., and Shannon, M. (2005). The serotonin syndrome. N. Engl. J. Med., 352 (11), 11121120. doi:10.1056/NEJMra041867Google Scholar
Breslau, N. (2002). Gender differences in trauma and posttraumatic stress disorder. J. Gend. Specif. Med., 5, 3440. doi:10.1037/0033–2909.132.6.959.959Google ScholarPubMed
Chandler, D. J., Gao, W. J., and Waterhouse, B. D. (2014). Heterogeneous organization of the locus coeruleus projections to prefrontal and motor cortices. Proc. Natl. Acad. Sci. U. S. A., 111(18), 68166821. doi:10.1073/pnas.1320827111Google Scholar
Corbetta, M., Patel, G., and Shulman, G. L. (2008). The reorienting system of the human brain: From environment to theory of mind. Neuron, 58, 306324. doi:10.1016/j.neuron.2008.04.017Google Scholar
Counts, S. E., and Mufson, E. J. (2012). Locus Coeruleus. In The Human Nervous System (pp. 425438). Elsevier Inc. doi:10.1016/B978-0-12-374236-0.10012-4Google Scholar
Datta, S., and MacLean, R. R. (2007). Neurobiological mechanisms for the regulation of mammalian sleep-wake behavior: Reinterpretation of historical evidence and inclusion of contemporary cellular and molecular evidence. Neurosci. Biobehav. Rev., 31, 775824. doi:10.1016/j.neubiorev.2007.02.004CrossRefGoogle ScholarPubMed
Davern, P. J. (2014). A role for the lateral parabrachial nucleus in cardiovascular function and fluid homeostasis. Front. Physiol., Published online 5(436). doi:10.3389/fphys.2014.00436Google Scholar
Delgado, P. L. (2000). Depression: the case for a monoamine deficiency. J. Clin. Psychiatry, 61(Suppl. 6), 711. Retrieved from www.ncbi.nlm.nih.gov/pubmed/10775018Google Scholar
Devilbiss, D. M., and Berridge, C. W. (2006). Low-dose methylphenidate actions on tonic and phasic locus coeruleus discharge. J. Pharmacol. Exp. Ther., 319, 13271335. doi:10.1124/jpet.106.110015Google Scholar
Ehrminger, M., Latimier, A., Pyatigorskaya, N., Garcia-Lorenzo, D., Leu-Semenescu, S., Vidaihet, M.,… Arnulf, I. (2016). The coeruleus/subcoeruleus complex in idiopathic rapid eye movement sleep behaviour disorder. Brain, 139(8), 11801188. doi:10.1093/brain/aww006Google Scholar
Ferrucci, M., Giorgi, F. S., Bartalucci, A., Busceti, C. L., and Fornai, F. (2013). The effects of locus coeruleus and norepinephrine in methamphetamine toxicity. Curr. Neuropharmacol., 11(1), 8094. doi:10.2174/157015913804999522Google Scholar
Guiard, B. P., El Mansari, M., and Blier, P. (2008). Cross-talk between dopaminergic and noradrenergic systems in the rat ventral tegmental area, locus coeruleus, and dorsal hippocampus. Mol. Pharmacol., 74, 14631475. doi:10.1124/mol.108.048033Google Scholar
Heneka, M. T., Nadrigny, F., Regen, T., Martinez-Hernandez, A., Dumitrescu-Ozimek, L., Terwil, D.Kummer, M. P. (2010). Locus ceruleus controls Alzheimer’s disease pathology by modulating microglial functions through norepinephrine. Proc. Nat. Acad. Sci. U. S. A., 107(13), 60586063. doi:10.1073/pnas.0909586107Google Scholar
Hinz, M., Stein, A., and Uncini, T. (2012). The discrediting of the monoamine hypothesis. Int. J. Gen. Med., 5, 135142. doi:10.2147/IJGM.S27824CrossRefGoogle ScholarPubMed
Holstege, G. (2009). The rostromedial tegmental nucleus and the emotional motor system: Role in basic survival behavior. J. Comp. Neurol., 513, 559565. doi:10.1002/cne.21990CrossRefGoogle ScholarPubMed
Holstege, G., Bandler, R., and Saper, C. B. (1996). The emotional motor system. Prog. Brain Res., 107, 36. Retrieved from www.ncbi.nlm.nih.gov/pubmed/8782510Google Scholar
Hornung, J-P. (2012). Raphe nuclei. In Mai, J. K., and Paxinos, G. (Eds.), The Human Nervous System (3rd edn., pp. 401424). New York, NY: Elsevier.Google Scholar
Jacobs, B. L. (1990). Locus coeruleus neuronal activity in behaving animals. In Heal, D. J. and Marsden, C. A. (Eds.), The Pharmacology of Noradrenaline in the Central Nervous System. (pp. 248265). Oxford, England: Oxford Medical Press.Google Scholar
Jacobs, B. L. (2002). Adult brain neurogenesis and depression. Brain Behav. Immunity, 16(5), 602609. doi:10.1016/S0889-1591(02)00015–6CrossRefGoogle ScholarPubMed
Jacobs, B. L., and Fornal, C. (1997). Serotonin and motor activity. Curr. Opin. Neurobiol., 7(6), 820825. doi:10.1016/S0959-4388(97)80141–9Google Scholar
Kawasaki, Y., Kumamoto, E., Furue, H., and Yoshimura, M. (2003). α2-Adrenoceptor-mediated presynaptic inhibition of primary afferent glutamatergic transmission in rat substantia gelatinosa neurons. Anesthesiology, 98, 682689. doi:10.1097/00000542–200303000-00016Google Scholar
Kessler, R. C. (2003). Epidemiology of women and depression. J. Affect. Disord., 74, 513. doi:10.1016/S0165-0327(02)00426–3Google Scholar
Kline, P., and Oertel, J. (1997). Depression associated with pontine vascular malformation. Biol. Psychiatry, 42, 519521. doi:10.1016/s0006-3223(97)00318–1Google Scholar
Kline, R. L., Zhang, S., Farr, O. M., Hu, S., Zaborszky, L., Samanez-Larkin, G. R., and Li, C-S. R. (2016). The effects of methylphenidate on resting-state functional connectivity of the basal nucleus of Meynert, locus coeruleus, and ventral tegmental area in healthy adults. Front. Hum. Neurosci., 18. doi:10.3389/fnhum.2016.00149Google Scholar
Kornstein, S. G., Schatzberg, A. F., Thase, M. E., Yonkers, K. A., McCullough, J. P.,Keitner, G. I.,… Keller, M. B. (2000). Gender differences in treatment response to sertraline versus imipramine in chronic depression. Am. J. Psychiatry, 157, 14451452. doi:10.1176/appi.ajp.157.9.1445Google Scholar
Kroeger, D., and de Lecea, L. (2009). The hypocretins and their role in narcolepsy. CNS Neurol. Disord. Drug Targets, 8(4), 271280. doi:10.2174/187152709788921645Google Scholar
Krueger, J. M., Frank, M. G., Wisor, J. P., and Roy, S. (2016). Sleep function: toward elucidating an enigma. Sleep Med. Rev., 28, 4250. doi:10.1016/j.smrv.2015.08.005Google Scholar
Kukkonen, J. P. (2012). Recent progress in orexin/hypocretin physiology and pharmacology. Biomol. Concepts, 3(5), 447463. doi:10.1515/bmc-2012–0013CrossRefGoogle ScholarPubMed
Lee, A., Wissekerke, A. E., Rosin, D. L., and Lynch, K, R. (1998). Localization of alpha2Cadrenergic receptor immunoreactivity in catecholaminergic neurons in the rat central nervous system. Neuroscience, 84, 10851096. Retrieved from www.ncbi.nlm.nih.gov/pubmed/9578397CrossRefGoogle ScholarPubMed
Li, C. S., Chung, S., Lu, D. P., and Cho, Y. K. (2012). Descending projections from the nucleus accumbens shell suppress activity of taste-responsive neurons in the hamster parabrachial nuclei. J. Neurophysiol., 108(5), 12881298. doi:10.1152/jn.00121.2012CrossRefGoogle ScholarPubMed
Lozano, A. M., Snyder, B. J., Hamani, C., Hutchison, W. D., and Dostrovsky, J. O. (2010). Basal ganglia physiology and deep brain stimulation. Mov. Disord., 25(suppl 1), S71–S75. doi:10.1002/mds.22714Google Scholar
Lumb, B. M. (2004) Hypothalamic and midbrain circuitry that distinguishes between escapable and inescapable pain. News Physiol. Sci., 19, 2226. doi:10.1152/nips.01467.2003Google Scholar
Lydic, R., and Baghdoyan, H. A. (2008) Acetylcholine modulates sleep and wakefulness: a synaptic perspective. In Monti, J. M., Pandi-Perumal, S. R., and Sinton, C. M. (Eds.). Neurochemistry of Sleep and Wakefulness. (pp. 109143). Cambridge, England: Cambridge University Press.Google Scholar
Mahlios, J., De la Herrán-Arita, A. K., and Mignot, E. (2013). The autoimmune basis of narcolepsy. Curr. Opin. Neurobiol., 23(5), 767773. doi:10.1016/j.conb.2013.04.013CrossRefGoogle ScholarPubMed
Maloney, K. J., Mainville, L., and Jones, B. E. (2002). c-Fos expression in dopaminergic and GABAergic neurons of the ventral mesencephalic tegmentum after paradoxical sleep deprivation and recovery. Eur. J. Neurosci., 15, 774778. doi:10.1046/j.1460–9568.2002.01907.xGoogle Scholar
Mansari, M. E., Guirard, B. P., Chernoloz, O., Ghanbari, R., Katz, N., and Blier, P. (2010). Relevance of norepinephrine-dopamine interactions in treatment of major depressive disorder. CNS Neurosci. Ther., 16(3), e1e17. doi:10.1111/j.1755–5949.2010.00146.xGoogle Scholar
McGinty, D. and Szymusiak, R. (2001). Brain structures and mechanisms involved in the generation of NREM sleep: focus on the preoptic hypothalamus. Sleep Med. Rev., 5, 323–316. doi:10.1053/smrv.2001.0170Google Scholar
Mendez, M. F., (1992). Pavor nocturnus from a brainstem glioma. J. Neurol. Neurosurg. Psychiatry, 55, 860. doi:10.1136/jnnp.55.9.860CrossRefGoogle ScholarPubMed
Mendez, M. F. and Bronstein, Y. L. (1999). Crying spells presenting as a transient ischemic attack. J. Neurol. Neurosurg. Psychiatry, 67, 255. doi:10.1136/jnnp.67.2.255Google Scholar
Miguelez, C., Morera-Herreras, T., Torrecilla, M., Ruiz-Ortega, J. A., and Ugedo, L. (2014). Interaction between the 5-HT system and the basal ganglia: Functional implication and therapeutic perspective in Parkinson’s disease. Front. Neural Circuits, 8, 21. doi:10.3389/fncir.2014.00021Google Scholar
Moruzzi, G., and Magoun, H. W. (1949). Brain stem reticular formation and activation of the EEG. Electroencephalography Clin. Neurophysiol., 1, 455473. Retrieved from www.ncbi.nlm.nih.gov/pubmed/7626974Google Scholar
Müller, M. L. T. M., and Bohnen, N. I. (2013). Cholinergic dysfunction in Parkinson’s disease. Curr. Neurol. Neurosci. Rep., 13, 377. doi:10.1007/s11910-013–0377-9CrossRefGoogle ScholarPubMed
Nakamura, K. (2013). The role of the dorsal raphé nucleus in reward-seeking behavior. Front. Integr. Neurosci., 7, 60. doi:10.3389/fnint.2013.00060Google Scholar
Nestler, E. J., and Aghajanian, G. K. (1997). Molecular and cellular basis of addiction. Science, 278, 5863. Retrieved from www.ncbi.nlm.nih.gov/pubmed/9311927Google Scholar
Nybäck, H. V., Walters, J. R., Aghajanian, G. K., and Roth, R. H. (1975).Tricyclic antidepressants: effects on the firing rate of brain noradrenergic neurons. Eur. J. Pharmacol., 32(2), 302312. Retrieved from www.ncbi.nlm.nih.gov/pubmed/1149813Google Scholar
Ordway, G. A., Schenk, J., Stockmeier, C. A., May, W., and Klimek, V. (2003). Elevated agonist binding to alpha2-adrenoceptors in the locus coeruleus in major depression. Biol. Psychiatry, 53(4), 315323. doi:10.1016/S0006-3223(02)01728–6Google Scholar
Perrier, J. F. (2016). Modulation of motoneuron activity by serotonin. Dan. Med. J., 63(2), pii: B5204. Abstract retrieved from www.ncbi.nlm.nih.gov/pubmed/26836802Google Scholar
Pertovaara, A. (2006). Noradrenergic pain modulation. Prog. Neurobiol., 80(2), 5358. doi:10.1016/j.pneurobio.2006.08.001Google Scholar
Puig, M. V., and Gulledge, A. T. (2011). Serotonin and prefrontal cortex function: Neurons, networks, and circuits. Mol. Neurobiol., 44(3). 449464. doi:10.1007/s12035-011–8214-0Google Scholar
Rajkowski, J., Majczynski, H., Clayton, E., and Aston-Jones, G. (2004). Activation of monkey locus ceruleus neurons varies with difficulty and behavioral performance in a target detection task. J. Neurophysiol., 92, 361371. doi:10.1152/jn.00673.2003Google Scholar
Reed, M. C., Nijhout, H. F., and Best, J. (2013). Computational studies of the role of serotonin in the basal ganglia. Front. Integr. Neurosci., May 24. doi:10.3389/fnint.2013.00041Google Scholar
Saha, S. (2005). Role of the central nucleus of the amygdala in the control of blood pressure: descending pathways to medullary cardiovascular nuclei. Clin. Exp. Pharmacol. Physiol., 32(5–6), 450456. doi:10.1111/j.1440–1681.2005.04210.xGoogle Scholar
Salomon, R. M., and Cowan, R. L. (2013). Oscillatory serotonin function in depression. Synapse, 67(11), 801820. doi:10.1002/syn.21675Google Scholar
Saper, C. B., Chou, T. C., and Scammell, T. E. (2001). The sleep switch: hypothalamic control of sleep and wakefulness. Trends Neurosci., 24(12), 726731. doi:10.1016/S0166-2236(00)02002–6Google Scholar
Savitz, J., Lucki, I., and Drevets, W. C. (2009). 5-HT1A receptor function in major depressive disorder. Prog. Neurobiol., 88(1), 1731. doi:10.1016/j.pneurobio.2009.01.009Google Scholar
Šimic, G., Babic, M., Leko, M. J., Wray, S., Harrington, C. R., Delalle, I.,… Hof, P. R. (2016). Monoaminergic neuropathology in Alzheimer’s disease. Prog. Neurobiol., 151, 131138. doi:10.1016/j.pneurobio.2016.04.001Google Scholar
Steininger, T. L., Gong, H., McGinty, D., and Szymusiak, R. (2001). Subregional organization of preoptic area/anterior hypothalamic projections to arousal-related monoaminergic cell groups. J. Comp. Neurol., 429, 638653. doi:10.1002/1096–9861(20010122)429:4<638::aid-cne10>3.3.CO;2-PGoogle Scholar
Stenberg, D. (2007). Neuroanatomy and neurochemistry of sleep. Cell. Mol. Life Sci., 64, 11871204. doi:10.1007/s00018-007–6530-3Google Scholar
Tully, K., and Bolshakov, V. Y. (2010). Emotional enhancement of memory: how norepinephrine enables synaptic plasticity. Mol. Brain, 3, 15. doi:10.1186/1756–6606-3–15Google Scholar
Vijayraghavan, S., Wang, M., Birnbaum, S. G., Williams, G. V., and Arnsten, A. F. (2007). Inverted-U dopamine D1 receptor actions on prefrontal neurons engaged in working memory. Nat. Neurosci., 10(3), 376384. doi: 10.1038/nn1846Google Scholar
Williams, J. T., Christie, M. J., and Manszoni, O. (2001). Cellular and synaptic adaptations mediating opioid dependence. Physiol. Rev., 81, 299343. Retrieved from www.ncbi.nlm.nih.gov/pubmed/11152760Google Scholar
Wu, Q., Boyle, M. P., and Palmiter, R. D. (2009). Loss of GABAergic signaling by AgRP neurons to the parabrachial nucleus leads to starvation. Cell, 137(7), 12251234. doi:10.1016/j.cell.2009.04.022.Google Scholar
Xing, B., Li, Y. C., and Gao, W. J. (2016). Norepinephrine versus dopamine and their interaction in modulating synaptic function in the prefrontal cortex. Bran Res., 1641(B), 217233. doi:10.1016/j.brainres.2016.01.005Google Scholar
Zangen, A., Nakash, R., Overstreet, D. H., and Yadid, G. (2001). Association between depressive behavior and absence of serotonin-dopamine interaction in the nucleus accumbens. Psychopharmacol., 155(4), 434439. doi:10.1007/s002130100746Google Scholar
Zhao, Z. Q., Chiechio, S., Sun, Y. G., Zhang, K. H., Zhao, C. S., Scott, M.,… Chen, Z-F. (2007). Mice lacking central serotonergic neurons show enhanced inflammatory pain and an impaired analgesic response to antidepressant drugs. J. Neurosci., 27(22), 60456053. doi:10.1523/JNEUROSCI.1623-07.2007Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • Brainstem
  • David L. Clark, Ohio State University, Nash N. Boutros, University of Missouri, Kansas City, Mario F. Mendez, University of California, Los Angeles
  • Book: The Brain and Behavior
  • Online publication: 22 February 2018
  • Chapter DOI: https://doi.org/10.1017/9781108164320.011
Available formats
×

Save book to Dropbox

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

  • Brainstem
  • David L. Clark, Ohio State University, Nash N. Boutros, University of Missouri, Kansas City, Mario F. Mendez, University of California, Los Angeles
  • Book: The Brain and Behavior
  • Online publication: 22 February 2018
  • Chapter DOI: https://doi.org/10.1017/9781108164320.011
Available formats
×

Save book to Google Drive

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

  • Brainstem
  • David L. Clark, Ohio State University, Nash N. Boutros, University of Missouri, Kansas City, Mario F. Mendez, University of California, Los Angeles
  • Book: The Brain and Behavior
  • Online publication: 22 February 2018
  • Chapter DOI: https://doi.org/10.1017/9781108164320.011
Available formats
×