Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-rkxrd Total loading time: 0 Render date: 2024-07-18T17:42:08.438Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  10 August 2009

Wilfrid Jänig
Affiliation:
Christian-Albrechts Universität zu Kiel, Germany
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
Integrative Action of the Autonomic Nervous System
Neurobiology of Homeostasis
, pp. 519 - 599
Publisher: Cambridge University Press
Print publication year: 2006

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

Ádám, G. (1967). Interoception and Behavior: An Experimental Study. Tanslated from the Hungarian by R. de Chantel, revised by H. Slucki. Budapest: Akadémiai Kiadó.Google Scholar
Ádám, G. (1998). Visceral Perception: Understanding Internal Cognition. New York: Plenum Press.CrossRefGoogle Scholar
Adams, D. J. and Harper, A. A. (1995). Electrophysiological properties of autonomic ganglia. In The Autonomic Nervous System, Vol. 6, Autonomic Ganglia, ed. McLachlan, E. M.. Chur, Switzerland: Harwood Academic Publisher, pp. 153–212.Google Scholar
Ader, A. and Cohen, N. (1993). Psychoneuroendocrinology: conditioning and stress. Annu. Rev. Physiol., 44, 53–85.Google Scholar
Aggleton, J. P. (ed.) (2000). The Amygdala: A Functional Analysis. Oxford: Oxford University Press.Google Scholar
Aicher, S. A., Kurucz, O. S., Reis, D. J. and Milner, T. A. (1995). Nucleus tractus solitarius efferent terminals synapse on neurons in the caudal ventrolateral medulla that project to the rostral ventrolateral medulla. Brain Res., 693, 51–63.CrossRefGoogle ScholarPubMed
Akasu, T. and Koketsu, K. (1986). Muscarinic transmission. In Autonomic and Enteric Ganglia, ed. Karczmar, K., Koketsu, K. and Nishi, S.. New York, London: Plenum Press, pp. 161–180.CrossRefGoogle Scholar
Akasu, T. and Nishimura, T. (1995). Synaptic transmission and function of parasympathetic ganglia. Prog. Neurobiol., 45, 459–522.CrossRefGoogle ScholarPubMed
Akert, K. (1981). Biological Order in Brain Organization. Selected Works of W. R. Hess. Berlin, Heidelberg, New York: Springer-Verlag.CrossRefGoogle Scholar
Akoev, G. N. (1981). Catecholamines, acetylcholine and excitability of mechanoreceptors. Prog. Neurobiol., 15, 269–294.CrossRefGoogle Scholar
Al Chaer, E. D., Lawand, N. B., Westlund, K. N. and Willis, W. D. (1996a). Visceral nociceptive input into the ventral posterolateral nucleus of the thalamus: a new function for the dorsal column pathway. J. Neurophysiol., 76, 2661–2674.CrossRefGoogle Scholar
Al Chaer, E. D., Lawand, N. B., Westlund, K. N. and Willis, W. D. (1996b). Pelvic visceral input into the nucleus gracilis is largely mediated by the postsynaptic dorsal column pathway. J. Neurophysiol., 76, 2675–2690.CrossRefGoogle Scholar
Al Chaer, E. D., Westlund, K. N. and Willis, W. D. (1997). Nucleus gracilis: an integrator for visceral and somatic information. J. Neurophysiol., 78, 521–527.CrossRefGoogle ScholarPubMed
Al Chaer, E. D., Feng, Y. and Willis, W. D. (1999). Comparative study of viscerosomatic input onto postsynaptic dorsal column and spinothalamic tract neurons in the primate. J. Neurophysiol., 82, 1876–1882.CrossRefGoogle ScholarPubMed
Alexander, R. S. (1946). Tonic and reflex functions of medullary sympathetic cardiovascular centers. J. Neurophysiol., 9, 205–217.CrossRefGoogle ScholarPubMed
Alexander, S. P. H., Mathie, A. and Peters, J. A. (2004). Guide to receptors and channels, 1st edition. Br. J. Pharmacol., 141 Suppl1, S1–S126.Google ScholarPubMed
Altschuler, S. M., Bao, X. M., Bieger, D., Hopkins, D. A. and Miselis, R. R. (1989). Viscerotopic representation of the upper alimentary tract in the rat: sensory ganglia and nuclei of the solitary and spinal trigeminal tracts. J. Comp. Neurol., 283, 248–268.CrossRefGoogle ScholarPubMed
Altschuler, S. M., Ferenci, D. A., Lynn, R. B. and Miselis, R. R. (1991). Representation of the cecum in the lateral dorsal motor nucleus of the vagus nerve and commissural subnucleus of the nucleus tractus solitarii in rat. J. Comp. Neurol., 304, 261–274.CrossRefGoogle ScholarPubMed
Altschuler, S. M., Rinaman, L. and Miselis, R. R. (1992). Viscerotopic representation of the alimentary tract in the dorsal and ventral vagal complexes in the rat. In Neuroanatomy and Physiology of Abdominal Vagal Afferents, ed. Ritter, S., Ritter, R. C. and Barnes, C. D.. Boca Raton: CRC Press, pp. 22–53.Google Scholar
Amann, R., Dray, A. and Hankins, M. W. (1988). Stimulation of afferent fibres of the guinea-pig ureter evokes potentials in inferior mesenteric ganglion neurones. J. Physiol., 402, 543–553.CrossRefGoogle ScholarPubMed
Amendt, K., Czachurski, J., Dembowsky, K. and Seller, H. (1979). Bulbospinal projections to the intermediolateral cell column: a neuroanatomical study. J. Auton. Nerv. Syst., 1, 103–107.CrossRefGoogle ScholarPubMed
An, X., Bandler, R., Öngür, D. and Price, J. L. (1998). Prefrontal cortical projections to longitudinal columns in the midbrain periaqueductal gray in macaque monkeys. J. Comp. Neurol., 401, 455–479.3.0.CO;2-6>CrossRefGoogle ScholarPubMed
Anders, S., Lotze, M., Erb, M., Grodd, W. and Birbaumer, N. (2004). Brain activity underlying emotional valence and arousal: a response-related fMRI study. Hum. Brain Mapp., 23, 200–209.CrossRefGoogle ScholarPubMed
Anderson, C. R., McAllen, R. M. and Edwards, S. L. (1995). Nitric oxide synthase and chemical coding in cat sympathetic postganglionic neurons. Neuroscience, 68, 255–264.CrossRefGoogle ScholarPubMed
Anderson, D. J. (1993). Molecular control of cell fate in the neural crest: the sympathoadrenal lineage. Annu. Rev. Neurosci., 16, 129–158.CrossRefGoogle ScholarPubMed
Anderson, E. A., Wallin, B. G. and Mark, A. L. (1987). Dissociation of sympathetic nerve activity in arm and leg muscle during mental stress. Hypertension, 9, III, 114–119.CrossRefGoogle ScholarPubMed
Anderson, R. L., Gibbins, I. L. and Morris, J. L. (1996). Non-noradrenergic sympathetic neurons project to extramuscular feed arteries and proximal intramuscular arteries of skeletal muscles in guinea-pig hindlimbs. J. Auton. Nerv. Syst., 61, 51–60.CrossRefGoogle ScholarPubMed
Andersson, K. E. (2001). Pharmacology of penile erection. Pharmacol. Rev., 53, 417–450.Google ScholarPubMed
Andreev, N. Y., Dimitrieva, N., Koltzenburg, M. and McMahon, S. B. (1995). Peripheral administration of nerve growth factor in the adult rat produces a thermal hyperalgesia that requires the presence of sympathetic post-ganglionic neurones. Pain, 63, 109–115.CrossRefGoogle ScholarPubMed
Andrew, D. and Craig, A. D. (2001a). Spinothalamic lamina I neurones selectively responsive to cutaneous warming in cats. J. Physiol., 537, 489–495.CrossRefGoogle Scholar
Andrew, D. and Craig, A. D. (2001b). Spinothalamic lamina I neurons selectively sensitive to histamine: a central neural pathway for itch. Nat. Neurosci., 4, 72–77.CrossRefGoogle Scholar
Andrew, D. and Craig, A. D. (2002a). Quantitative responses of spinothalamic lamina I neurones to graded mechanical stimulation in the cat. J. Physiol., 545, 913–931.CrossRefGoogle Scholar
Andrew, D. and Craig, A. D. (2002b). Responses of spinothalamic lamina I neurons to maintained noxious mechanical stimulation in the cat. J. Neurophysiol., 87, 1889–1901.CrossRefGoogle Scholar
Andrews, P. L. R. (1986). Vagal afferent innervation of the gastrointestinal tract. Prog. Brain Res., 67, 65–86.CrossRefGoogle ScholarPubMed
Anrep, G. V., Pascual, W. and Rössler, R. (1936a). Respiratory variations of the heart rate. I. The reflex mechanism of the sinus arrhythmia. Proc. Royal Soc. Lond., 119 (Series B), 191–217.CrossRefGoogle Scholar
Anrep, G. V., Pascual, W. and Rössler, R. (1936b). Respiratory variations of the heart rate. II. The central mechanism of the sinus arrhythmia and the inter-relationships between central and reflex mechanism. Proc. Royal Soc. Lond., 119 (Series B), 218–230.CrossRefGoogle Scholar
Apkarian, A. V. and Shi, T. (1994). Squirrel monkey lateral thalamus. I. Somatic nociresponsive neurons and their relation to spinothalamic terminals. J. Neurosci., 14, 6779–6795.CrossRefGoogle ScholarPubMed
Apodaca, G. (2004). The uroepithelium: not just a passive barrier. Traffic, 5, 117–128.CrossRefGoogle Scholar
Apodaca, G., Kiss, S., Ruiz, W.et al. (2003). Disruption of bladder epithelium barrier function after spinal cord injury. Am. J. Physiol. Renal Physiol., 284, F966–F976.CrossRefGoogle ScholarPubMed
Appenzeller, O. and Oribe, E. (1997). The Autonomic Nervous System. An Introduction to Basic and Clinical Concepts, 5th edn. Amsterdam: Elsevier.Google Scholar
Appenzeller, O. (ed.) (1999). Handbook of Clinical Neurology, Vol. 74, The Autonomic Nervous System, part I: Normal Functions. Amsterdam: Elsevier.Google Scholar
Appenzeller, O. (ed.) (2000). Handbook of Clinical Neurology, Vol. 75, The Autonomic Nervous System, part II: Dysfunctions. Amsterdam: Elsevier.Google Scholar
Applegate, C. D., Kapp, B. S., Underwood, M. D. and McNall, C. L. (1983). Autonomic and somatomotor effects of amygdala central N. stimulation in awake rabbits. Physiol. Behav., 31, 353–360.CrossRefGoogle ScholarPubMed
Araki, I. (1994). Inhibitory postsynaptic currents and the effects of γ-aminobutyric acid on visually identified sacral parasympathetic preganglionic neurons in neonatal rats. J. Neurophysiol., 72, 2903–2910.CrossRefGoogle ScholarPubMed
Araki, I. and Groat, W. C. (1996). Unitary excitatory synaptic currents in preganglionic neurons mediated by two distinct groups of interneurons in neonatal rat sacral parasympathetic nucleus. J. Neurophysiol., 76, 215–226.CrossRefGoogle ScholarPubMed
Araki, I. and Groat, W. C. (1997). Developmental synaptic depression underlying reorganization of visceral reflex pathways in the spinal cord. J. Neurosci., 17, 8402–8407.CrossRefGoogle ScholarPubMed
Bachoo, M. and Polosa, C. (1985). Properties of a sympatho-inhibitory and vasodilator reflex evoked by superior laryngeal nerve afferents in the cat. J. Physiol., 364, 183–198.CrossRefGoogle ScholarPubMed
Bachoo, M. and Polosa, C. (1986). The pattern of sympathetic neurone activity during expiration in the cat. J. Physiol., 378, 375–390.CrossRefGoogle ScholarPubMed
Bachoo, M. and Polosa, C. (1987). Properties of the inspiration-related activity of sympathetic preganglionic neurones of the cervical trunk in the cat. J. Physiol., 385, 545–564.CrossRefGoogle ScholarPubMed
Bacon, S. J., Zagon, A. and Smith, A. D. (1990). Electron microscopic evidence of a monosynaptic pathway between cells in the caudal raphe nuclei and sympathetic preganglionic neurons in the rat spinal cord. Exp. Brain Res., 79, 589–602.CrossRefGoogle ScholarPubMed
Bacon, S. J. and Smith, A. D. (1993). A monosynaptic pathway from an identified vasomotor centre in the medial prefrontal cortex to an autonomic area in the thoracic spinal cord. Neuroscience, 54, 719–728.CrossRefGoogle Scholar
Badoer, E. (2001). Hypothalamic paraventricular nucleus and cardiovascular regulation. Clin. Exp. Pharmacol. Physiol., 28, 95–99.CrossRefGoogle ScholarPubMed
Bahns, E., Ernsberger, U., Jänig, W. and Nelke, A. (1986a). Functional characteristics of lumbar visceral afferent fibres from the urinary bladder and the urethra in the cat. Pflügers Arch., 407, 510–518.CrossRefGoogle Scholar
Bahns, E., Ernsberger, U., Jänig, W. and Nelke, A. (1986b). Discharge properties of mechanosensitive afferents supplying the retroperitoneal space. Pflügers Arch., 407, 519–525.CrossRefGoogle Scholar
Bahns, E., Halsband, U. and Jänig, W. (1987). Responses of sacral visceral afferents from the lower urinary tract, colon and anus to mechanical stimulation. Pflügers Arch., 410, 296–303.CrossRefGoogle ScholarPubMed
Bahr, R., Blumberg, H. and Jänig, W. (1981). Do dichotomizing afferent fibers exist which supply visceral organs as well as somatic structures? A contribution to the problem of referred pain. Neurosci. Lett., 24, 25–28.CrossRefGoogle ScholarPubMed
Bahr, R., Bartel, B., Blumberg, H. and Jänig, W. (1986a). Functional characterization of preganglionic neurons projecting in the lumbar splanchnic nerves: neurons regulating motility. J. Auton. Nerv. Syst., 15, 109–130.CrossRefGoogle Scholar
Bahr, R., Bartel, B., Blumberg, H. and Jänig, W. (1986b). Functional characterization of preganglionic neurons projecting in the lumbar splanchnic nerves: vasoconstrictor neurons. J. Auton. Nerv. Syst., 15, 131–140.CrossRefGoogle Scholar
Bahr, R., Bartel, B., Blumberg, H. and Jänig, W. (1986c). Secondary functional properties of lumbar visceral preganglionic neurons. J. Auton. Nerv. Syst., 15, 141–152.CrossRefGoogle Scholar
Bainton, C. R., Richter, D. W., Seller, H., Ballantyne, D. and Klein, J. P. (1985). Respiratory modulation of sympathetic activity. J. Auton. Nerv. Syst., 12, 77–90.CrossRefGoogle ScholarPubMed
Baker, D. G., Coleridge, H. M., Coleridge, J. C. and Nerdrum, T. (1980). Search for a cardiac nociceptor: stimulation by bradykinin of sympathetic afferent nerve endings in the heart of the cat. J. Physiol., 306, 519–536.CrossRefGoogle ScholarPubMed
Baldissera, F., Hultborn, H. and Illert, M. (1981). Integration in spinal neuronal systems. In Handbook of Physiology, Section 1, The Nervous System, Vol. IIb, Motor Control, part I. ed. Brooks, V. B.. Bethesda: American Physiological Society, pp. 509–595.Google Scholar
Baldwin, B. A., Parrott, R. F. and Ebenezer, I. S. (1998). Food for thought: a critique on the hypothesis that endogenous cholecystokinin acts as a physiological satiety factor. Prog. Neurobiol., 55, 477–507.CrossRefGoogle ScholarPubMed
Bamshad, M., Aoki, V. T., Adkison, M. G., Warren, W. S. and Bartness, T. J. (1998). Central nervous system origins of the sympathetic nervous system outflow to white adipose tissue. Am. J. Physiol., 275, R291–R299.Google ScholarPubMed
Bamshad, M., Song, C. K. and Bartness, T. J. (1999). central nervous system origins of the sympathetic nervous system outflow to brown adipose tissue. Am. J. Physiol., 276, R1569–R1578.Google ScholarPubMed
Bandler, R. (1988). Brain mechanisms of aggression as revealed by electrical and chemical stimulation: suggestion of a central role for the midbrain periaqueductal grey region. Prog. Psychobiol. Physiol. Psychol., 13, 67–153.Google Scholar
Bandler, R. and Shipley, M. T. (1994). Columnar organization in the midbrain periaqueductal gray: modules for emotional expression?Trends Neurosci., 17, 379–389.CrossRefGoogle ScholarPubMed
Bandler, R. and Keay, K. A. (1996). Columnar organization in the midbrain periaqueductal gray and the integration of emotional expression. Prog. Brain Res., 107, 285–300.CrossRefGoogle ScholarPubMed
Bandler, R., Carrive, P. and Zhang, S. P. (1991). Integration of somatic and autonomic reactions within the midbrain periaqueductal grey: viscerotopic, somatotopic and functional organization. Prog. Brain Res., 87, 269–305.CrossRefGoogle ScholarPubMed
Bandler, R., Keay, K. A., Floyd, N. and Price, J. (2000a). Central circuits mediating patterned autonomic activity during active vs. passive emotional coping. Brain Res. Bull., 53, 95–104.CrossRefGoogle Scholar
Bandler, R., Price, J. L. and Keay, K. A. (2000b). Brain mediation of active and passive emotional coping. Prog. Brain Res., 122, 333–349.CrossRefGoogle Scholar
Bao, J. X., Gonon, F. and Stjärne, L. (1993). Frequency- and train length-dependent variation in the roles of postjunctional alpha 1- and alpha 2-adrenoceptors for the field stimulation-induced neurogenic contraction of rat tail artery. Naunyn-Schmiedeberg's Arch. Pharmacol., 347, 601–616.CrossRefGoogle ScholarPubMed
Barcroft, H., Brod, Z., Hejl, Z., Hirsjärvi, E. A. and Kitchen, A. H. (1960). The mechanism of the vasodilatation in the forearm muscle during stress (mental arithmetic). Clin. Sci., 19, 577–586.Google Scholar
Bard, P. (1928). A diencephalic mechanism for the expression of rage with special reference to the sympathetic nervous system. Am. J. Physiol., 84, 490–515.Google Scholar
Bard, P. (1932). An emotional expression after decortication with some remarks on certain theoretical views. Part I. Psychol. Rev., 41, 309–329.CrossRefGoogle Scholar
Bard, P. (1960). Anatomical organization of the central nervous system in relation to control of the heart and blood vessels. Physiol. Rev., 40 (Suppl4), 3–26.Google Scholar
Bard, P. and Macht, M. B. (1957). The behavior of chronically decerebrate cats. In Ciba Foundation Symposium on the Neurological Basis of Behavior, ed. Wolstenholme, G. E. W. and O'Connor, M.. Boston: Little, Brown and Co., pp. 55–71.Google Scholar
Bard, P. and Rioch, D. McK. (1937). A study of four cats deprived of neocortex and additional portions of the forebrain. Bull. Johns Hopkins Hosp., 60, 73–147.Google Scholar
Barker, D. and Saito, M. (1981). Autonomic innervation of receptors and muscle fibres in cat skeletal muscle. Proc. R. Soc. London Biol., 212, 317–332.CrossRefGoogle ScholarPubMed
Barman, S. M. and Gebber, G. L. (1976). Basis for synchronization of sympathetic and phrenic nerve discharges. Am. J. Physiol., 231, 1601–1607.Google ScholarPubMed
Barman, S. M. and Gebber, G. L. (1984). Spinal interneurons with sympathetic nerve-related activity. Am. J. Physiol., 247, R761–R767.Google ScholarPubMed
Barman, S. M. and Gebber, G. L. (2000). “Rapid” rhythmic discharges of sympathetic nerves: sources, mechanisms of generation, and physiological relevance. J. Biol. Rhythms, 15, 365–379.CrossRefGoogle ScholarPubMed
Barman, S. M., Orer, H. S. and Gebber, G. L. (2001). The role of the medullary lateral tegmental field in the generation and baroreceptor reflex control of sympathetic nerve discharge in the cat. Ann. N. Y. Acad. Sci., 940, 270–285.CrossRefGoogle ScholarPubMed
Barnes, P. J. (ed.) (1997). The Autonomic Nervous System, Vol. 7, Autonomic Control of the Respiratory System. Amsterdam: Harwood Academic Publishers.Google Scholar
Baron, R. and Jänig, W. (1988). Neurons projecting rostrally in the hypogastric nerve of the cat. J. Auton. Nerv. Syst., 24, 81–86.CrossRefGoogle ScholarPubMed
Baron, R. and Jänig, W. (1991). Afferent and sympathetic neurons projecting into lumbar visceral nerves of the male rat. J. Comp. Neurol., 314, 429–436.CrossRefGoogle ScholarPubMed
Baron, R., Jänig, W. and McLachlan, E. M. (1985a). On the anatomical organization of the lumbosacral sympathetic chain and the lumbar splanchnic nerves of the cat – Langley revisited. J. Auton. Nerv. Syst., 12, 289–300.CrossRefGoogle Scholar
Baron, R., Jänig, W. and McLachlan, E. M. (1985b). The afferent and sympathetic components of the lumbar spinal outflow to the colon and pelvic organs in the cat: I. The hypogastric nerve. J. Comp. Neurol., 238, 135–146.CrossRefGoogle Scholar
Baron, R., Jänig, W. and McLachlan, E. M. (1985c). The afferent and sympathetic components of the lumbar spinal outflow to the colon and pelvic organs in the cat. II. The lumbar splanchnic nerves. J. Comp. Neurol., 238, 147–157.CrossRefGoogle Scholar
Baron, R., Jänig, W. and McLachlan, E. M. (1985d). The afferent and sympathetic components of the lumbar spinal outflow to the colon and pelvic organs in the cat. III. The colonic nerves, incorporating an analysis of all components of the lumbar prevertebral outflow. J. Comp. Neurol., 238, 158–168.CrossRefGoogle Scholar
Baron, R., Jänig, W. and Kollmann, W. (1988). Sympathetic and afferent somata projecting in hindlimb nerves and the anatomical organization of the lumbar sympathetic nervous system of the rat. J. Comp. Neurol., 275, 460–468.CrossRefGoogle ScholarPubMed
Baron, R., Jänig, W. and With, H. (1995). Sympathetic and afferent neurones projecting into forelimb and trunk nerves and the anatomical organization of the thoracic sympathetic outflow of the rat. J. Auton. Nerv. Syst., 53, 205–214.CrossRefGoogle ScholarPubMed
Barraco, I. R. A. (ed.) (1994). Nucleus of the Solitary Tract. Boca Raton: CRC Press.Google Scholar
Bartel, B., Blumberg, H. and Jänig, W. (1986). Discharge patterns of motility-regulating neurons projecting in the lumbar splanchnic nerves to visceral stimuli in spinal cats. J. Auton. Nerv. Syst., 15, 153–163.CrossRefGoogle ScholarPubMed
Bartness, T. J. and Bamshad, M. (1998). Innervation of mammalian white adipose tissue: implications for the regulation of total body fat. Am. J. Physiol., 275, R1399–R1411.Google ScholarPubMed
Bartsch, T., Häbler, H. J. and Jänig, W. (1996). Functional properties of postganglionic sympathetic neurones supplying the submandibular gland in the anaesthetized rat. Neurosci. Lett., 214, 143–146.CrossRefGoogle ScholarPubMed
Bartsch, T., Häbler, H. J. and Jänig, W. (1999). Hypoventilation recruits preganglionic sympathetic fibers with inspiration-related activity in the superior cervical trunk of the rat. J. Auton. Nerv. Syst., 77, 31–38.CrossRefGoogle ScholarPubMed
Bartsch, T., Häbler, H. J. and Jänig, W. (2000). Reflex patterns of preganglionic sympathetic fibers projecting to the superior cervical ganglion in the rat. Auton. Neurosci., 83, 66–74.CrossRefGoogle ScholarPubMed
Baumann, T. K., Simone, D. A., Shain, C. N. and LaMotte, R. H. (1991). Neurogenic hyperalgesia: the search for the primary cutaneous afferent fibers that contribute to capsaicin-induced pain and hyperalgesia. J. Neurophysiol., 66, 212–227.CrossRefGoogle ScholarPubMed
Bayliss, W. M. (1901). On the origin from the spinal cord of the vaso-dilator fibres of the hind-limb, and on the nature of these fibres. J. Physiol., 26, 173–209.CrossRefGoogle ScholarPubMed
Bayliss, W. M. and Starling, E. H. (1899). The movements and innervation of the small intestine. J. Physiol., 24, 99–143.CrossRefGoogle ScholarPubMed
Bayliss, W. M. and Starling, E. H. (1900). The movements and innervation of the large intestine. J. Physiol., 26, 107–118.CrossRefGoogle ScholarPubMed
Beckstead, R. M., Morse, J. R. and Norgren, R. (1980). The nucleus of the solitary tract in the monkey: projections to the thalamus and brain stem nuclei. J. Comp. Neurol., 190, 259–282.CrossRefGoogle ScholarPubMed
Bell, C., Jänig, W., Kümmel, H. and Xu, H. (1985). Differentiation of vasodilator and sudomotor responses in the cat paw pad to preganglionic sympathetic stimulation. J. Physiol., 364, 93–104.CrossRefGoogle ScholarPubMed
Belmonte, C. and Cervero, F. (eds.) (1996). Neurobiology of Nociception. Oxford, New York, Tokyo: Oxford University Press.CrossRefGoogle Scholar
Benison, S., Barger, A. C. and Wolfe, E. L. (1987). Walter B. Cannon. The Life and Times of a Young Scientist. Cambridge Mass. London England: The Belknarp Press of Harvard University Press.Google Scholar
Bennett, T. and Gardiner, S. M. (eds.) (1996). The Autonomic Nervous System, Vol. 8, Nervous Control of Blood Vessels. Amsterdam: Harwood Academic Publishers.Google Scholar
Berkley, K. J., Robbins, A. and Sato, Y. (1988). Afferent fibers supplying the uterus in the rat. J. Neurophysiol., 59, 142–163.CrossRefGoogle ScholarPubMed
Berkley, K. J., Hotta, H., Robbins, A. and Sato, Y. (1990). Functional properties of afferent fibers supplying reproductive and other pelvic organs in pelvic nerve of female rat. J. Neurophysiol., 63, 256–272.CrossRefGoogle ScholarPubMed
Berkley, K. J., Robbins, A. and Sato, Y. (1993). Functional differences between afferent fibers in the hypogastric and pelvic nerves innervating female reproductive organs in the rat. J. Neurophysiol., 69, 533–544.CrossRefGoogle ScholarPubMed
Bernard, C. (1858). Leçons Sur la Physiologie et la Pathologie du Système Nerveau [On the Physiology and Pathophysiology of the Nervous System]. Paris: Bailliere.Google Scholar
Bernard, C. (1957) [1865]. Introduction à l'Étude de la Médicine Experimentale. [An Introduction to the Study of Experimental Medicine]. New York: Dover Publications. Paris: Baillière.Google Scholar
Bernard, C. (1974) [1878]. Lecons sur les Phénomènes de la Vie Communes aux Animaux et aux Végétaux [Lectures on the Phenomena of Life Common to Animals and Plants]. Translated by H. E. Hoff, R. Guillemin and L. Guillemin. Paris (Springfield Illinois): B. Ballière et Fils (Thomas).Google Scholar
Bernard, J. F. and Bandler, R. (1998). Parallel circuits for emotional coping behaviour: new pieces in the puzzle. J. Comp. Neurol., 401, 429–436.3.0.CO;2-3>CrossRefGoogle ScholarPubMed
Bernard, J. F. and Besson, J. M. (1990). The spino(trigemino)pontoamygdaloid pathway: electrophysiological evidence for an involvement in pain processes. J. Neurophysiol., 63, 473–490.CrossRefGoogle ScholarPubMed
Bernard, J. F., Huang, G. F. and Besson, J. M. (1994). The parabrachial area: electrophysiological evidence for an involvement in visceral nociceptive processes. J. Neurophysiol., 71, 1646–1660.CrossRefGoogle ScholarPubMed
Berne, C. and Fagius, J. (1986). Skin sympathetic activity during insulin-induced hypoglycemia. Diabetologia, 29, 855–860.CrossRefGoogle Scholar
Berntson, G. G. and Cacioppo, J. T. (2000). From homeostasis to allodynamic regulation. In Handbook of Psychophysiology, 2nd edn., eds. Cacioppo, J. T., Tassinary, L. G. and Berntson, G. G.. Cambridge: Cambridge University Press, pp. 459–481.Google Scholar
Berthoud, H. R. and Neuhuber, W. L. (2000). Functional and chemical anatomy of the afferent vagal system. Auton. Neurosci., 85, 1–17.CrossRefGoogle ScholarPubMed
Berthoud, H. R. and Powley, T. L. (1990). Identification of vagal preganglionics that mediate cephalic phase insulin response. Am. J. Physiol., 258, R523–R530.Google ScholarPubMed
Berthoud, H. R. and Powley, T. L. (1992). Vagal afferent innervation of the rat fundic stomach: morphological characterization of the gastric tension receptor. J. Comp. Neurol., 319, 261–276.CrossRefGoogle ScholarPubMed
Berthoud, H. R., Fox, E. A. and Powley, T. L. (1990). Localization of vagal preganglionics that stimulate insulin and glucagon secretion. Am. J. Physiol., 258, R160–R168.Google ScholarPubMed
Berthoud, H. R., Carlson, N. R. and Powley, T. L. (1991). Topography of efferent vagal innervation of the rat gastrointestinal tract. Am. J. Physiol., 260, R200–R207.Google ScholarPubMed
Berthoud, H. R., Patterson, L. M., Willing, A. E., Mueller, K. and Neuhuber, W. L. (1997). Capsaicin-resistant vagal afferent fibers in the rat gastrointestinal tract: anatomical identification and functional integrity. Brain Res., 746, 195–206.CrossRefGoogle ScholarPubMed
Besedovsky, H. O. and del Rey, A. (1992). Immune-neuroendocrine circuits: integrative role of cytokines. Front. Neuroendocrinol., 13, 61–94.Google ScholarPubMed
Besedovsky, H. O. and del Rey, A. (1995). Immune-neuroendocrine interactions: facts and hypotheses. Endocr. Rev., 17, 64–102.CrossRefGoogle Scholar
Bester, H., Chapman, V., Besson, J. M. and Bernard, J. F. (2000). Physiological properties of the lamina I spinoparabrachial neurons in the rat. J. Neurophysiol., 83, 2239–2259.CrossRefGoogle ScholarPubMed
Beyak, M. J. and Grundy, D. (2005). Vagal afferents innervating the gastrointestinal tract. In Advance in Vagal Afferent Neurobiology, ed. Undem, B. and Weinreich, D.. Boca Raton: CRC Press, pp. 315–350.CrossRefGoogle Scholar
Bianchi, A. L., Denavit-Saubie, M. and Champagnat, J. (1995). Central control of breathing in mammals: neuronal circuitry, membrane properties, and neurotransmitters. Physiol. Rev., 75, 1–45.CrossRefGoogle ScholarPubMed
Bieger, D. and Hopkins, D. A. (1987). Viscerotopic representation of the upper alimentary tract in the medulla oblongata in the rat: the nucleus ambiguus. J. Comp. Neurol., 262, 546–562.CrossRefGoogle ScholarPubMed
Bielefeld, T. K. and Gebhart, G. F. (2005). Visceral pain: basic mechanisms. In Wall and Melzack's Textbook of Pain, 5th edn., ed. McMahon, S. B. and Koltzenburg, M.. Amsterdam, Edinburgh: Elsevier Churchill Livingstone, pp. 721–736.Google Scholar
Bini, G., Hagbarth, K. E., Hynninen, P. and Wallin, B. G. (1980a). Thermoregulatory and rhythm-generating mechanisms governing the sudomotor and vasoconstrictor outflow in human cutaneous nerves. J. Physiol., 306, 547–552.CrossRefGoogle Scholar
Bini, G., Hagbarth, K. E., Hynninen, P. and Wallin, B. G. (1980b). Regional similarities and differences in thermoregulatory vaso- and sudomotor tone. J. Physiol., 306, 553–565.CrossRefGoogle Scholar
Bini, G., Hagbarth, K. E. and Wallin, B. G. (1981). Cardiac rhythmicity of skin sympathetic activity recorded from peripheral nerves in man. J. Auton. Nerv. Syst., 4, 17–24.CrossRefGoogle ScholarPubMed
Birder, L. A. (2005). More than just a barrier: urothelium as a drug target for urinary bladder pain. Am. J. Physiol. Renal. Physiol., 289, F489–F495.CrossRefGoogle ScholarPubMed
Björntop, P. (1997). Behavior and metabolic disease. Int. J. Behav. Med., 3, 285–302.CrossRefGoogle Scholar
Blackman, J. G. (1974). Function of autonomic ganglia. In The Peripheral Nervous System, ed. Hubbard, J. I.. New York: Plenum Press, pp. 257–276.CrossRefGoogle Scholar
Blackman, J. G., Ginsborg, B. L. and Ray, C. (1963). Synaptic transmission in the sympathetic ganglion of the frog. J. Physiol., 167, 355–373.CrossRefGoogle ScholarPubMed
Blair, D. A., Glover, W. E., Greenfield, A. D. M. and Roddie, I. C. (1959). Excitation of cholinergic vasodilator nerves to human skeletal muscles during emotional stress. J. Physiol., 148, 633–646.CrossRefGoogle ScholarPubMed
Blessing, W. W. (1997). The Lower Brain Stem and Bodily Homeostasis. New York, Oxford: Oxford University Press.Google Scholar
Blessing, W. W. and Nalivaiko, E. (2000). Regional blood flow and nociceptive stimuli in rabbits: patterning by medullary raphe, not ventrolateral medulla. J. Physiol., 524, 279–292.CrossRefGoogle Scholar
Blessing, W. W., Li, Y. W. and Wesselingh, S. L. (1991). Transneuronal transport of herpes simplex virus from the cervical vagus to brain neurons with axonal inputs to central vagal sensory nuclei in the rat. Neuroscience, 42, 261–274.CrossRefGoogle ScholarPubMed
Blessing, W. W., Yu, Y. H. and Nalivaiko, E. (1999). Raphe pallidus and parapyramidal neurons regulate ear pinna vascular conductance in the rabbit. Neurosci. Lett., 270, 33–36.CrossRefGoogle ScholarPubMed
Blix, A. S. and Folkow, B. (1983). Cardiovascular adjustments to diving in mammals and birds. In Handbook of Physiology, Section 2: The Cardiovascular System Vol. III: Peripheral Circulation, ed. Shepherd, J. T. and Abboud, F. M.. Bethesda: American Physiological Society, pp. 917–945.Google Scholar
Blok, B. F. and Holstege, G. (1996). The neuronal control of micturition and its relation to the emotional motor system. Prog. Brain Res., 107, 113–126.CrossRefGoogle ScholarPubMed
Blumberg, H. and Jänig, W. (1982). Changes in unmyelinated fibers including sympathetic postganglionic fibers of a skin nerve after peripheral neuroma formation. J. Auton. Nerv. Syst., 6, 173–183.CrossRefGoogle ScholarPubMed
Blumberg, H. and Jänig, W. (1983a). Enhancement of resting activity in postganglionic vasoconstrictor neurones following short-lasting repetitive activation of preganglionic axons. Pflügers Arch., 396, 89–94.CrossRefGoogle Scholar
Blumberg, H. and Jänig, W. (1983b). Changes of reflexes in vasoconstrictor neurons supplying the cat hindlimb following chronic nerve lesions: a model for studying mechanisms of reflex sympathetic dystrophy? J. Auton. Nerv. Syst., 7, 399–411.CrossRefGoogle Scholar
Blumberg, H. and Jänig, W. (1985). Reflex patterns in postganglionic vasoconstrictor neurons following chronic nerve lesions. J. Auton. Nerv. Syst., 14, 157–180.CrossRefGoogle ScholarPubMed
Blumberg, H. and Wallin, B. G. (1987). Direct evidence of neurally mediated vasodilatation in hairy skin of the human foot. J. Physiol., 382, 105–121.CrossRefGoogle ScholarPubMed
Blumberg, H., Jänig, W., Rieckmann, C. and Szulczyk, P. (1980). Baroreceptor and chemoreceptor reflexes in postganglionic neurones supplying skeletal muscle and hairy skin. J. Auton. Nerv. Syst., 2, 223–240.CrossRefGoogle ScholarPubMed
Blumberg, H., Haupt, P., Jänig, W. and Kohler, W. (1983). Encoding of visceral noxious stimuli in the discharge patterns of visceral afferent fibres from the colon. Pflügers Arch., 398, 33–40.CrossRefGoogle ScholarPubMed
Boczek-Funcke, A., Häbler, H. J., Jänig, W. and Michaelis, M. (1991). Rapid phasic baroreceptor inhibition of the activity in sympathetic preganglionic neurones does not change throughout the respiratory cycle. J. Auton. Nerv. Syst., 34, 185–194.CrossRefGoogle Scholar
Boczek-Funcke, A., Dembowsky, K., Häbler, H. J.et al. (1992a). Classification of preganglionic neurones projecting into the cat cervical sympathetic trunk. J. Physiol., 453, 319–339.CrossRefGoogle Scholar
Boczek-Funcke, A., Dembowsky, K., Häbler, H. J., Jänig, W. and Michaelis, M. (1992b). Respiratory-related activity patterns in preganglionic neurones projecting into the cat cervical sympathetic trunk. J. Physiol., 457, 277–296.CrossRefGoogle Scholar
Boczek-Funcke, A., Häbler, H. J., Jänig, W. and Michaelis, M. (1992c). Respiratory modulation of the activity in sympathetic neurones supplying muscle, skin and pelvic organs in the cat. J. Physiol., 449, 333–361.CrossRefGoogle Scholar
Boczek-Funcke, A., Dembowsky, K., Häbler, H. J., Jänig, W. and Michaelis, M. (1993). Spontaneous activity, conduction velocity and segmental origin of different classes of thoracic preganglionic neurons projecting into the cat cervical sympathetic trunk. J. Auton. Nerv. Syst., 43, 189–200.CrossRefGoogle ScholarPubMed
Bogduk, N. (1983). The innervation of the lumbar spine. Spine, 8, 286–293.CrossRefGoogle ScholarPubMed
Bogduk, N., Windsor, M. and Inglis, A. (1988). The innervation of the cervical intervertebral discs. Spine, 13, 2–8.CrossRefGoogle ScholarPubMed
Bolme, P. and Fuxe, K. (1970). Adrenergic and cholinergic nerve terminals in skeletal muscle vessels. Acta Physiol. Scand., 78, 52–59.CrossRefGoogle ScholarPubMed
Bolme, B., Novotny, J., Uvnäs, B. and Wright, P. G. (1970). Species distribution of sympathetic cholinergic vasodilator nerves in skeletal muscle. Acta Physiol. Scand., 78, 60–64.CrossRefGoogle ScholarPubMed
Bolter, C. P., Wallace, D. J. and Hirst, G. D. (2001). Failure of Ba2 + and Cs+ to block the effects of vagal nerve stimulation in sinoatrial node cells of the guinea-pig heart. Auton. Neurosci., 94, 93–101.CrossRefGoogle ScholarPubMed
Bolton, B., Carmichael, E. A. and Stürup, G. (1936). Vaso-constriction following deep inspiration. J. Physiol., 86, 83–94.CrossRefGoogle ScholarPubMed
Bors, E. H. and Comarr, A. E. (1960). Neurological disturbances of sexual function with special reference to 529 patients with spinal cord injury. Urol. Survey, 10, 191–222.Google Scholar
Bos, J. D. (ed.) (1989). Skin Immune System. Boca Raton: CRC Press.Google Scholar
Bos, J. D. and Kapsenberg, M. L. (1986). The skin immune system. Its cellular constituents and their interactions. Immunol. Today, 7, 235–240.CrossRefGoogle ScholarPubMed
Boscan, P. and Paton, J. F. (2002). Integration of cornea and cardiorespiratory afferents in the nucleus of the solitary tract of the rat. Am. J. Physiol. Heart Circ. Physiol., 282, H1278–H1287.CrossRefGoogle ScholarPubMed
Boscan, P., Kasparov, S. and Paton, J. F. (2002). Somatic nociception activates neurokinin 1 receptors in the nucleus tractus solitarii to attenuate the baroreceptor cardiac reflex. Eur. J. Neurosci., 16, 907–920.CrossRefGoogle ScholarPubMed
Bosnjak, Z. J. and Kampine, J. P. (1982). Intracellular recordings from the stellate ganglion of the cat. J. Physiol., 324, 273–283.CrossRefGoogle ScholarPubMed
Bosnjak, Z. J. and Kampine, J. P. (1984). Peripheral neural input to neurons of the middle cervical ganglion in the cat. Am. J. Physiol., 246, R354–R358.Google ScholarPubMed
Bosnjak, Z. J. and Kampine, J. P. (1985). Electrophysiological and morphological characterization of neurons in stellate ganglion of cats. Am. J. Physiol., 248, R288–R292.Google ScholarPubMed
Boucsein, W. (1992). Electrodermal Activity. New York: Plenum Press.CrossRefGoogle Scholar
Boyd, H. D., McLachlan, E. M., Keast, J. R. and Inokuchi, H. (1996). Three electrophysiological classes of guinea pig sympathetic postganglionic neurone have distinct morphologies. J. Comp. Neurol., 369, 372–387.3.0.CO;2-2>CrossRefGoogle ScholarPubMed
Bramich, N. J., Edwards, F. R. and Hirst, G. D. (1990). Sympathetic nerve stimulation and applied transmitters on the sinus venosus of the toad. J. Physiol., 429, 349–375.CrossRefGoogle ScholarPubMed
Bramich, N. J., Brock, J. A., Edwards, F. R. and Hirst, G. D. (1993). Responses to sympathetic nerve stimulation of the sinus venosus of the toad. J. Physiol., 461, 403–430.CrossRefGoogle ScholarPubMed
Bramich, N. J., Brock, J. A., Edwards, F. R. and Hirst, G. D. (1994). Ionophoretically applied acetylcholine and vagal stimulation in the arrested sinus venosus of the toad, Bufo marinus. J. Physiol., 478, 289–300.Google ScholarPubMed
Bramich, N. J., Cousins, H. M., Edwards, F. R. and Hirst, G. D. (2001). Parallel metabotropic pathways in the heart of the toad, Bufo marinus. Am. J. Physiol. Heart Circ. Physiol., 281, H1771–H1777.CrossRefGoogle ScholarPubMed
Brock, J. A. and Cunnane, T. C. (1988). Electrical activity at the sympathetic neuroeffector junction in the guinea-pig vas deferens. J. Physiol., 399, 607–632.CrossRefGoogle ScholarPubMed
Brock, J. A. and Cunnane, T. C. (1992). Impulse conduction in sympathetic nerve terminals in the guinea-pig vas deferens and the role of the pelvic ganglia. Neuroscience, 47, 185–196.CrossRefGoogle ScholarPubMed
Brock, J. A. and Cunnane, T. C. (1993). Neurotransmitter release mechanisms at the sympathetic neuroeffector junction. Exp. Physiol., 78, 591–614.CrossRefGoogle ScholarPubMed
Brock, J. A. and Helden, D. F. (1995). Enhanced excitatory junction potentials in mesenteric arteries from spontaneously hypertensive rats. Pflügers Arch., 430, 901–908.CrossRefGoogle ScholarPubMed
Brock, J. A., McLachlan, E. M. and Rayner, S. E. (1997). Contribution of alpha-adrenoceptors to depolarization and contraction evoked by continuous asynchronous sympathetic nerve activity in rat tail artery. Br. J. Pharmacol., 120, 1513–1521.CrossRefGoogle ScholarPubMed
Brodal, P. (1998). The Central Nervous System. Structure and Function. New York, Oxford: Oxford University Press.Google Scholar
Brooke, R. E., Pyner, S., McLeish, P.et al. (2002). Spinal cord interneurones labelled transneuronally from the adrenal gland by a GFP-herpes virus construct contain the potassium channel subunit Kv3.1b. Auton. Neurosci., 98, 45–50.CrossRefGoogle ScholarPubMed
Brooke, R. E., Deuchars, J. and Deuchars, S. A. (2004). Input-specific modulation of neurotransmitter release in the lateral horn of the spinal cord via adenosine receptors. J. Neurosci., 24, 127–137.CrossRefGoogle ScholarPubMed
Brookes, S. J. (2001). Classes of enteric nerve cells in the guinea-pig small intestine. Anat. Rec., 262, 58–70.3.0.CO;2-V>CrossRefGoogle ScholarPubMed
Brookes, S. and Costa, M. (eds.) (2002). The Autonomic Nervous System, Vol. 14, Innervation of the Gastrointestinal Tract. London, New York: Francis and Taylor.Google Scholar
Brooks, C. M., Koizumi, K. and Pinkston, J. O. (eds.) (1975). The Life and Contributions of Walter Bradford Cannon 1871–1945. New York: State University of New York, Downstate Medical Center.Google Scholar
Brown, A. M. (1967). Cardiac sympathetic adrenergic pathways in which synaptic transmission is blocked by atropine sulfate. J. Physiol., 191, 271–288.CrossRefGoogle ScholarPubMed
Brown, A. M. (1969). Sympathetic ganglionic transmission and the cardiovascular changes of the defense reaction in the cat. Circ. Res., 24, 843–849.CrossRefGoogle ScholarPubMed
Brown, A., Ricci, M. J. and Weaver, L. C. (2004). nerve growth factor message and protein distribution in the injured rat spinal cord. Exp. Neurol., 188, 115–127.CrossRefGoogle ScholarPubMed
Brown, D. L. and Guyenet, P. G. (1984). Cardiovascular neurons of brain stem with projections to spinal cord. Am. J. Physiol., 247, R1009–R1016.Google ScholarPubMed
Brown, D. L. and Guyenet, P. G. (1985). Electrophysiological study of cardiovascular neurons in the rostral ventrolateral medulla in rats. Circ. Res., 56, 359–369.CrossRefGoogle ScholarPubMed
Browning, K. N., Renehan, W. E. and Travagli, R. A. (1999). Electrophysiological and morphological heterogeneity of rat dorsal vagal neurones which project to specific areas of the gastrointestinal tract. J. Physiol., 517, 521–532.CrossRefGoogle ScholarPubMed
Browning, K. N., Coleman, F. H. and Travagli, R. (2005). Effects of pancreatic polypeptide on pancreas-projecting rat dorsal motor nucleus of the vagus neuron. Am. J. Physiol. Gastrointest. Liver Physiol., 289, G209–G219.CrossRefGoogle Scholar
Bruce, A. N. (1910). Über die Beziehung der sensiblen Nervenendigungen zum Entzündungsvorgang [On the relation between sensory nerve endings and inflammation]. Arch. Exptl. Pathol. Pharmakol., 63, 424–433.CrossRefGoogle Scholar
Bruce, A. N. (1913). Vaso-dilator axon-reflexes. Q. J. Exp. Physiol., 6, 339–354.CrossRefGoogle Scholar
Burnett, A. L., Lowenstein, C. J., Bredt, D. S., Chang, T. S. and Snyder, S. H. (1992). Nitric oxide: a physiologic mediator of penile erection. Science, 257, 401–403.CrossRefGoogle ScholarPubMed
Burns, A. J., Lomax, A. E., Torihashi, S., Sanders, K. M. and Ward, S. M. (1996). Interstitial cells of Cajal mediate inhibitory neurotransmission in the stomach. Proc. Natl. Acad. Sci. U.S.A., 93, 12008–12013.CrossRefGoogle ScholarPubMed
Burnstock, G. and Hoyle, C. H. V. (eds.) (1992). The Autonomic Nervous System, Vol. 1, Autonomic Neuroeffector Mechanisms. Chur, Switzerland: Harwood Academic Publishers.Google Scholar
Burnstock, G. and Sillito, A. M. (eds.) (2000). The Autonomic Nervous System, Vol. 13, Nervous Control of the Eye. Amsterdam: Harwood Academic Publishers.Google Scholar
Busse, R., Edwards, G., Feletou, M.et al. (2002). EDHF: bringing the concepts together. Trends Pharmacol. Sci., 23, 374–380.CrossRefGoogle ScholarPubMed
Butler, P. J. and Jones, D. R. (1997). Physiology of diving of birds and mammals. Physiol. Rev., 77, 837–899.CrossRefGoogle ScholarPubMed
Bykov, K. M. (1959) [1944]. The Cerebral Cortex and the Internal Organs. [Translated from Russian and edited by R. Hodes and A. Kilbey.] Moscow: Foreign Language Publishing House.Google Scholar
Cabot, J. B. (1990). Sympathetic preganglionic neurons: cytoarchitecture, ultrastructure, and biophysical properties. In Central Regulation of Autonomic Functions, eds. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 44–67.Google Scholar
Cabot, J. B. (1996). Some principles of the spinal organization of the sympathetic preganglionic outflow. Prog. Brain Res., 107, 29–42.CrossRefGoogle ScholarPubMed
Cabot, J. B., Alessi, V., Carroll, J. and Ligorio, M. (1994). Spinal cord lamina V and lamina VII interneuronal projections to sympathetic preganglionic neurons. J. Comp. Neurol., 347, 515–530.CrossRefGoogle ScholarPubMed
Cajal, S. R. (1995) [1911]. Histologie du Système Nerveux de l'Homme et des Vertèbrès. [Histology of the Nervous System of Man and Vertebrates]. Maloine, Vol. 2, edited and translated by Swanson, L. W. and Swanson, N.. Oxford: Oxford University Press.Google Scholar
Campbell, G. D., Edwards, F. R., Hirst, G. D. S. and O'Shea, J. E. (1989). Effects of vagal stimulation and applied acetylcholine on pacemaker potentials in the guinea-pig heart. J. Physiol., 415, 57–68.CrossRefGoogle ScholarPubMed
Campos, R. R. and McAllen, R. M. (1997). Cardiac sympathetic premotor neurons. Am. J. Physiol., 272, R615–R620.Google ScholarPubMed
Campos, R. R. and McAllen, R. M. (1999). Tonic drive to sympathetic premotor neurons of rostral ventrolateral medulla from caudal pressor area neurons. Am. J. Physiol., 276, R1209–R1213.Google ScholarPubMed
Canning, B. J. and Mazzone, S. B. (2005). Reflexes initiated by activation of the vagal afferent nerves innervating the airways and lung. In Advances in Vagal Afferent Neurobiology, ed. Undem, B. J. and Weinreich, D., Boca Raton: CRC, Taylor & Francis, pp. 403–430.CrossRefGoogle Scholar
Cannon, W. B. (1911). The Mechanical Factors of Digestion. London: Edward Arnold.Google Scholar
Cannon, W. B. (1914a). The interrelations of emotions as suggested by recent physiological researches. Am. J. Physiol., 25, 252–282.Google Scholar
Cannon, W. B. (1914b). The emergency function of the adrenal medulla in pain and the major emotions. Am. J. Physiol., 33, 356–372.Google Scholar
Cannon, W. B. (1927). The James-Lange theory of emotions: a critical examination and an alternative theory. Am. J. Psychol., 39, 106–124.CrossRefGoogle Scholar
Cannon, W. B. (1928). Die Notfallfunktion des sympathico-adrenalen Systems [The emergency function of the sympathico-adrenal system]. Ergebn. Physiol., 27, 380–406.CrossRefGoogle Scholar
Cannon, W. B. (1929a). Organization for physiological homeostasis. Physiol. Rev., 9, 399–431.CrossRefGoogle Scholar
Cannon, W. B. (1929b). Bodily Changes in Pain, Hunger, Fear and Rage. New York: Appleton.Google Scholar
Cannon, W. B. (1933). A method of stimulating autonomic nerves in the anesthetized cat with observation on the motor and sensory effects. Am. J. Physiol., 105, 366–372.Google Scholar
Cannon, W. B. (1939). The Wisdom of the Body, 2nd revised and enlarged edition. New York: Norton.Google Scholar
Cannon, W. B. and Murphy, F. T. (1906). The movements of the stomach and intestine in some surgical conditions. Ann. Surg., 43, 512–536.CrossRefGoogle ScholarPubMed
Cannon, W. B., Newton, H. F., Bright, E. M., Menkin, V. and Moore, R. M. (1929). Some aspects of the physiology of animals surviving complete exclusion of sympathetic nerve impulses. Am. J. Physiol., 89, 84–107.Google Scholar
Cano, G., Card, J. P., Rinaman, L. and Sved, A. F. (2000). Connections of Barrington's nucleus to the sympathetic nervous system in rats. J. Auton. Nerv. Syst., 79, 117–128.CrossRefGoogle ScholarPubMed
Cano, G., Sved, A. F., Rinaman, L., Rabin, B. S. and Card, J. P. (2001). Characterization of the central nervous system innervation of the rat spleen using viral transneuronal tracing. J. Comp. Neurol., 439, 1–18.CrossRefGoogle ScholarPubMed
Card, J. P., Rinaman, L., Schwaber, J. S.et al. (1990). Neurotropic properties of pseudorabies virus: uptake and transneuronal passage in the rat central nervous system. J. Neurosci., 10, 1974–1994.CrossRefGoogle ScholarPubMed
Card, J. P., Swanson, L. W. and Moore, R. Y. (2003). The hypothalamus: an overview of regulatory systems. In Fundamental Neuroscience, 2nd edn., eds. Squire, L. R., Bloom, F. E., McConnell, S. K., et al. San Diego: Academic Press, pp. 897–909.Google Scholar
Carrive, P. (1993). The periaqueductal gray and defensive behavior: functional representation and neuronal organization. Behav. Brain Res., 58, 27–47.CrossRefGoogle ScholarPubMed
Carrive, P. and Morgan, M. M. (2004). Periaqueductal grey. In The Human Nervous System, ed. Paxinos, G. and Mai, J. K.. Amsterdam: Elsevier Academic Press, pp. 393–423.Google Scholar
Cassell, J. F. and McLachlan, E. M. (1986). The effect of a transient outward current (IA) on synaptic potentials in sympathetic ganglion cells of the guinea-pig. J. Physiol., 374, 273–288.CrossRefGoogle ScholarPubMed
Cassell, J. F., Clark, A. L. and McLachlan, E. M. (1986). Characteristics of phasic and tonic sympathetic ganglion cells of the guinea-pig. J. Physiol., 372, 457–483.CrossRefGoogle ScholarPubMed
Cassell, J. F., McLachlan, E. M. and Sittiracha, T. (1988). The effect of temperature on neuromuscular transmission in the main caudal artery of the rat. J. Physiol., 397, 31–49.CrossRefGoogle ScholarPubMed
Casson, D. M. and Ronald, K. (1975). The harp seal, Pagophilus groenlandicus (Erxleben, 1777). XIV. Cardiac arrythmias. Comp. Biochem. Physiol. A, 50, 307–314.CrossRefGoogle ScholarPubMed
Causing, C. G., Gloster, A., Aloyz, R.et al. (1997). Synaptic innervation density is regulated by neuron-derived brain-derived neurotrophic factor. Neuron, 18, 257–267.CrossRefGoogle Scholar
Cechetto, D. F. (1995). Supraspinal mechanisms of visceral pain. In Visceral Pain, ed. Gebhardt, G. F.. Seattle: IASP Press, pp. 261–290.Google Scholar
Cechetto, D. F. and Saper, C. B. (1990). Role of the cerebral cortex in autonomic function. In Central Regulation of Autonomic Functions, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 208–223.Google Scholar
Celander, O. (1954). The range of control exercised by the sympatho-adrenal system. Acta Physiol. Scand. Suppl., 116, 1–132.Google Scholar
Cervero, F. (1982). Afferent activity evoked by natural stimulation of the biliary system in the ferret. Pain, 13, 137–151.CrossRefGoogle ScholarPubMed
Cervero, F. (1994). Sensory innervation of the viscera: peripheral basis of visceral pain. Physiol. Rev., 74, 95–138.CrossRefGoogle ScholarPubMed
Cervero, F. (1995). Mechanisms of visceral pain: past and present. In Visceral Pain, ed. Gebhart, G. F.. Seattle: IASP Press, pp. 25–41.Google Scholar
Cervero, F. (1996). Visceral nociceptors. In Neurobiology of Nociceptors, ed. Belmonte, C. and Cervero, F.. Oxford, New York, Toronto: Oxford University Press, pp. 220–240.CrossRefGoogle Scholar
Cervero, F. and Connell, L. A. (1984). Distribution of somatic and visceral primary afferent fibers within the thoracic spinal cord of the cat. J. Comp. Neurol., 230, 88–98.CrossRefGoogle ScholarPubMed
Cervero, F. and Jänig, W. (1992). Visceral nociceptors: a new world order? Trends Neurosci., 15, 374–378.CrossRefGoogle ScholarPubMed
Cervero, F. and Morrison, J. F. B. (eds.) (1986). Visceral Sensation. Prog. Brain Res. Vol. 67, Amsterdam: Elsevier.Google Scholar
Cervero, F. and Sann, H. (1989). Mechanically evoked responses of afferent fibers innervating the guinea-pig's ureter: an in vitro study. J. Physiol., 412, 245–266.CrossRefGoogle Scholar
Cervero, F. and Tattersall, J. E. (1986). Somatic and visceral sensory integration in the thoracic spinal cord. Prog. Brain Res., 67, 189–205.CrossRefGoogle ScholarPubMed
Chan, R. K. and Sawchenko, P. E. (1998). Organization and transmitter specificity of medullary neurons activated by sustained hypertension: implications for understanding baroreceptor reflex circuitry. J. Neurosci., 18, 371–387.CrossRefGoogle ScholarPubMed
Chandler, M. J., Zhang, J., Qin, C. and Foreman, R. D. (2002). Spinal inhibitory effects of cardiopulmonary afferent inputs in monkeys: neuronal processing in high cervical segments. J. Neurophysiol., 87, 1290–1302.CrossRefGoogle ScholarPubMed
Chang, H. S., Staras, K. and Gilbey, M. P. (2000). Multiple oscillators provide metastability in rhythm generation. J. Neurosci., 20, 5135–5143.CrossRefGoogle ScholarPubMed
Chang, H. S., Staras, K., Smith, J. E. and Gilbey, M. P. (1999). Sympathetic neuronal oscillators are capable of dynamic synchronization. J. Neurosci., 19, 3183–3197.CrossRefGoogle ScholarPubMed
Chapleau, M. W. and Abboud, F. (eds.) (2001). Neuro-Cardiovascular Regulation: from Molecules to Man. New York: The New York Academy of Sciences.Google Scholar
Chau, D. and Schramm, L. P. (1997). Sympathetically correlated activity of dorsal horn neurons in spinally transected rats. J. Neurophysiol., 77, 2966–2974.CrossRefGoogle ScholarPubMed
Chau, D., Johns, D. G. and Schramm, L. P. (2000). Ongoing and stimulus-evoked activity of sympathetically correlated neurons in the intermediate zone and dorsal horn of acutely spinalized rats. J. Neurophysiol., 83, 2699–2707.CrossRefGoogle ScholarPubMed
Cheng, Z. and Powley, T. L. (2000). Nucleus ambiguus projections to cardiac ganglia of rat atria: an anterograde tracing study. J. Comp. Neurol., 424, 588–606.3.0.CO;2-7>CrossRefGoogle Scholar
Cheng, Z., Powley, T. L., Schwaber, J. S. and Doyle, F. J., III. (1999). Projections of the dorsal motor nucleus of the vagus to cardiac ganglia of rat atria: an anterograde tracing study. J. Comp. Neurol., 410, 320–341.3.0.CO;2-5>CrossRefGoogle Scholar
Chizh, B. A., Headley, P. M. and Paton, J. F. (1998). Coupling of sympathetic and somatic motor outflows from the spinal cord in a perfused preparation of adult mouse in vitro. J. Physiol., 508, 907–918.CrossRefGoogle Scholar
Choate, J. K., Edwards, F. R., Hirst, G. D. and O'Shea, J. E. (1993a). Effects of sympathetic nerve stimulation on the sino-atrial node of the guinea-pig. J. Physiol., 471, 707–727.CrossRefGoogle Scholar
Choate, J. K., Klemm, M. and Hirst, G. D. S. (1993b). Sympathetic and parasympathetic neuromuscular junctions in the guinea-pig sino-atrial node. J. Auton. Nerv. Syst., 44, 1–15.CrossRefGoogle Scholar
Christensen, J. (1994). The motility of the colon. In Physiology of the Gastrointestinal Tract, ed. Johnson, L. R.. New York: Raven Press, pp. 991–1024.Google Scholar
Christian, E. P. and Weinreich, D. (1988). Long-duration spike afterhyperpolarizations in neurons from the guinea pig superior cervical ganglion. Neurosci. Lett., 84, 191–196.CrossRefGoogle ScholarPubMed
Chrousos, G. P. (1998). Stressors, stress, and neuroendocrine integration of the adaptive response. The 1997 Hans Selye Memorial Lecture. Ann. New York Acad. Sci., 851, 311–335.CrossRefGoogle Scholar
Ciriello, J., Hochstenbach, S. L. and Roder, S. (1994). Central projections of baroreceptor and chemoreceptor afferent fibers in the rat. In Nucleus of the Solitary Tract, ed. Barraco, I. R. A.. Boca Raton: CRC Press, pp. 35–50.Google Scholar
Clayton, E. C. and Williams, C. L. (2000). Adrenergic activation of the nucleus tractus solitarius potentiates amygdala norepinephrine release and enhances retention performance in emotionally arousing and spatial memory tasks. Behav. Brain Res., 112, 151–158.CrossRefGoogle ScholarPubMed
Clement, C. I., Keay, K. A., Owler, B. K. and Bandler, R. (1996). Common patterns of increased and decreased fos expression in midbrain and pons evoked by noxious deep somatic and noxious visceral manipulations in the rat. J. Comp. Neurol., 366, 495–515.3.0.CO;2-#>CrossRefGoogle ScholarPubMed
Clement, C. I., Keay, K. A., Podzebenko, K., Gordon, B. D. and Bandler, R. (2000). Spinal sources of noxious visceral and noxious deep somatic afferent drive onto the ventrolateral periaqueductal gray of the rat. J. Comp. Neurol., 425, 323–344.3.0.CO;2-Z>CrossRefGoogle ScholarPubMed
Clutter, W. E., Bier, D. M., Shah, S. D. and Cryer, P. E. (1980). Epinephrine plasma metabolic clearance rates and physiologic thresholds for metabolic and hemodynamic actions in man. J. Clin. Invest., 66, 94–101.CrossRefGoogle ScholarPubMed
Cochrane, K. L. and Nathan, M. A. (1989). Normotension in conscious rats after placement of bilateral electrolytic lesions in the rostral ventrolateral medulla. J Auton. Nerv. Syst., 26, 199–211.CrossRefGoogle ScholarPubMed
Cochrane, K. L. and Nathan, M. A. (1993). Cardiovascular effects of lesions of the rostral ventrolateral medulla and the nucleus reticularis parvocellularis in rats. J. Auton. Nerv. Syst., 43, 69–81.CrossRefGoogle ScholarPubMed
Coderre, T. J., Basbaum, A. I. and Levine, J. D. (1989). Neural control of vascular permeability: interaction between primary afferents, mast cells, and sympathetic efferents. J. Neurophysiol., 62, 48–58.CrossRefGoogle ScholarPubMed
Coleridge, H. M. and Coleridge, J. C. (1980). Cardiovascular afferents involved in regulation of peripheral vessels. Annu. Rev. Physiol., 42, 413–427.CrossRefGoogle ScholarPubMed
Coleridge, H. M., Coleridge, J. C. G. (1997). Afferent nerves in the airways. In Autonomic Control of the Respiratory System, ed. Barnes, P. J.. Amsterdam: Harwood Academic Publishers GmbH, pp. 39–58.Google Scholar
Coleridge, H. M., Coleridge, J. C., Kaufman, M. P. and Dangel, A. (1981). Operational sensitivity and acute resetting of aortic baroreceptors in dogs. Circ. Res., 48, 676–684.CrossRefGoogle ScholarPubMed
Coleridge, J. C. and Coleridge, H. M. (1984). Afferent vagal C fibre innervation of the lungs and airways and its functional significance. Rev. Physiol. Biochem. Pharmacol., 99, 1–110.CrossRefGoogle ScholarPubMed
Contreras, R. J., Gomez, M. M. and Norgren, R. (1980). Central origins of cranial nerve parasympathetic neurons in the rat. J. Comp. Neurol., 190, 373–394.CrossRefGoogle ScholarPubMed
Cooke, H. J. (1994). Neuroimmune signaling in regulation of intestinal ion transport. Am. J. Physiol., 266, G167–G178.Google ScholarPubMed
Cooke, H. J. (1998). “Enteric tears”: chloride secretion and its neural regulation. News Physiol. Sci., 13, 269–274.Google ScholarPubMed
Cooke, H. J. and Reddix, R. A. (1994). Neural regulation of intestinal electrolyte transport. In Physiology of the Gastrointestinal Tract, 3rd edn., ed. Johnson, L. R.. New York: Raven Press, pp. 2083–2132.Google Scholar
Coolen, L. M., Allard, J., Truit, W. A. and McKenna, K. E. (2004). Central regulation of ejaculation. Physiol. Behav., 83, 203–213.CrossRefGoogle ScholarPubMed
Coonan, E. M., Downie, J. W. and Du, H. J. (1999). Sacral spinal cord neurons responsive to bladder pelvic and perineal inputs in cats. Neurosci. Lett., 260, 137–140.CrossRefGoogle ScholarPubMed
Coote, J. H. (1984). Spinal and supraspinal reflex pathways of cardio-cardiac sympathetic reflexes. Neurosci. Lett., 46, 243–247.CrossRefGoogle ScholarPubMed
Coote, J. H. (1988). The organisation of cardiovascular neurons in the spinal cord. Rev. Physiol. Biochem. Pharmacol., 110, 147–285.CrossRefGoogle ScholarPubMed
Coote, J. H. and Downman, C. B. (1966). Central pathways of some autonomic reflex discharges. J. Physiol., 183, 714–729.CrossRefGoogle ScholarPubMed
Coote, J. H. and Sato, A. (1978). Supraspinal regulation of spinal reflex discharge into cardiac sympathetic nerves. Brain Res., 142, 425–437.CrossRefGoogle ScholarPubMed
Coote, J. H., Macleod, V. H., Fleetwood-Walker, S. M. and Gilbey, M. P. (1981). Baroreceptor inhibition of sympathetic activity at a spinal site. Brain Res., 220, 81–93.CrossRefGoogle Scholar
Costa, M., Brookes, S. J., Steele, P. A.et al. (1996). Neurochemical classification of myenteric neurons in the guinea-pig ileum. Neuroscience, 75, 949–967.CrossRefGoogle ScholarPubMed
Coupland, R. E. (1965). Electron microscopic observations on the structure of the rat adrenal medulla. I. The ultrastructure and organization of chromaffin cells in the normal adrenal medulla. J. Anat., 99, 231–254.Google ScholarPubMed
Cousins, H. M., Edwards, F. R., Hirst, G. D. and Wendt, I. R. (1993). Cholinergic neuromuscular transmission in the longitudinal muscle of the guinea-pig ileum. J. Physiol., 471, 61–86.CrossRefGoogle ScholarPubMed
Cousins, H. M., Edwards, F. R. and Hirst, G. D. (1995). Neuronally released and applied acetylcholine on the longitudinal muscle of the guinea-pig ileum. Neuroscience, 65, 193–207.CrossRefGoogle ScholarPubMed
Coutinho, S. V., Su, X., Sengupta, J. N. and Gebhart, G. F. (2000). Role of sensitized pelvic nerve afferents from the inflamed rat colon in the maintenance of visceral hyperalgesia. Prog. Brain Res., 129, 375–387.CrossRefGoogle Scholar
Cowley, A. W. Jr. (1992). Long-term control of arterial blood pressure. Physiol. Rev., 72, 231–300.CrossRefGoogle ScholarPubMed
Cowley, A. W. Jr., Liard, J. F. and Guyton, A. C. (1973). Role of the baroreceptor reflex in daily control of arterial blood pressure and other variables in dogs. Circ. Res., 32, 564–576.CrossRefGoogle ScholarPubMed
Cox, G. E., Jordan, D., Paton, J. F., Spyer, K. M. and Wood, L. M. (1987). Cardiovascular and phrenic nerve responses to stimulation of the amygdala central nucleus in the anaesthetized rabbit. J. Physiol., 389, 541–556.CrossRefGoogle ScholarPubMed
Craig, A. D. (1996). An ascending general homeostatic afferent pathway originating in lamina I. Prog. Brain Res., 107, 225–243.CrossRefGoogle ScholarPubMed
Craig, A. D. (2002). How do you feel? Interoception: the sense of the physiological condition of the body. Nat. Rev. Neurosci., 3, 655–666.CrossRefGoogle Scholar
Craig, A. D. (2003a). Pain mechanisms: labeled lines versus convergence in central processing. Annu. Rev. Neurosci., 26, 1–30.CrossRefGoogle Scholar
Craig, A. D. (2003b). A new view of pain as a homeostatic emotion. Trends Neurosci., 26, 303–307.CrossRefGoogle Scholar
Craig, A. D. (2003c). Interoception: the sense of the physiological condition of the body. Curr. Opin. Neurobiol., 13, 500–505.CrossRefGoogle Scholar
Craig, A. D. (2004a). Lamina I, but not lamina V, spinothalamic neurons exhibit responses that correspond with burning pain. J. Neurophysiol., 92, 2604–2609.CrossRefGoogle Scholar
Craig, A. D. (2004b). Distribution of trigeminothalamic and spinothalamic lamina I terminations in the macaque monkey. J. Comp. Neurol., 477, 119–148.CrossRefGoogle Scholar
Craig, A. D. and Blomqvist, A. (2002). Is there a specific lamina I spinothalamocortical pathway for pain and temperature sensations in primates?J. Pain, 3, 95–101.CrossRefGoogle Scholar
Craig, A. D. and Kniffki, K. D. (1985). Spinothalamic lumbosacral lamina I cells responsive to skin and muscle stimulation in the cat. J. Physiol., 365, 197–221.CrossRefGoogle ScholarPubMed
Craig, A. D. and Mense, S. (1983). The distribution of afferent fibers from the gastrocnemius-soleus muscle in the dorsal horn of the cat, as revealed by the transport of horseradish peroxidase. Neurosci. Lett., 41, 233–238.CrossRefGoogle ScholarPubMed
Craig, A. D., Heppelmann, B. and Schaible, H. G. (1988). The projection of the medial and posterior articular nerves of the cat's knee to the spinal cord. J. Comp. Neurol., 276, 279–288.CrossRefGoogle ScholarPubMed
Craig, A. D., Bushnell, M. C., Zhang, E. T. and Blomqvist, A. (1994). A thalamic nucleus specific for pain and temperature sensation. Nature, 372, 770–773.CrossRefGoogle ScholarPubMed
Craig, A. D., Krout, K. and Andrew, D. (2001). Quantitative response characteristics of thermoreceptive and nociceptive lamina I spinothalamic neurons in the cat. J. Neurophysiol., 86, 1459–1480.CrossRefGoogle ScholarPubMed
Cravo, S. L., Morrison, S. F. and Reis, D. J. (1991). Differentiation of two cardiovascular regions within caudal ventrolateral medulla. Am. J. Physiol., 261, R985–R994.Google ScholarPubMed
Crawford, J. P. and Frankel, H. L. (1971). Abdominal ‘visceral’ sensation in human tetraplegia. Paraplegia, 9, 153–158.Google ScholarPubMed
Critchley, H. and Dolan, R. J. (2004). Central representation of autonomic states. In Human Brain Function. 2nd edn., eds. Frackowiak, R. S. J., Friston, K. J., Frith, C. D.et al. Amsterdam: Elsevier Academic Press, pp. 397–417.Google Scholar
Crowcroft, P. J., Holman, M. E. and Szurszewski, J. H. (1971). Excitatory input from the distal colon to the inferior mesenteric ganglion in the guinea-pig. J. Physiol., 219, 443–461.CrossRefGoogle ScholarPubMed
Cryer, P. E. (1980). Physiology and pathophysiology of the human sympathoadrenal neuroendocrine system. New Engl. J. Med., 303, 436–444.Google ScholarPubMed
Cunnane, T. C. and Stjärne, L. (1982). Secretion of transmitter from individual varicosities of guinea-pig and mouse vas deferens: all-or-none and extremely intermittent. Neuroscience, 7, 2565–2576.CrossRefGoogle ScholarPubMed
Cyon, E. and Ludwig, C. (1866). Die Reflexe eines der sensiblen Nerven des Herzens auf die motorischen der Blutgefässe [The reflex of one of the heart nerves on the motor nerves to blood vessels]. Ber. Sächs. Ges. Wiss., 18, 307–328.Google Scholar
Czachurski, J., Dembowsky, K., Seller, H., Nobiling, R. and Taugner, R. (1988). Morphology of electrophysiologically identified baroreceptor afferents and second order neurones in the brainstem of the cat. Arch. Ital. Biol., 126, 129–144.Google ScholarPubMed
Dado, R. J., Katter, J. T. and Giesler, G. J. Jr. (1994). Spinothalamic and spinohypothalamic tract neurons in the cervical enlargement of rats. I. Locations of antidromically identified axons in the thalamus and hypothalamus. J. Neurophysiol., 71, 959–980.CrossRefGoogle Scholar
Dail, W. G. (1993). Autonomic innervation of male reproductive genitalia. In The Autonomic Nervous System, Vol. 3, Nervous Control of the Urogenital System, ed. Maggi, C. A.. Chur, Switzerland: Harwood Academic Publishers, pp. 69–101.Google Scholar
Dalsgaard, C. J. and Elfvin, L. G. (1982). Structural studies on the connectivity of the inferior mesenteric ganglion of the guinea pig. J. Auton. Nerv. Syst., 5, 265–278.CrossRefGoogle ScholarPubMed
Dalsgaard, C. J., Hökfelt, T., Elfvin, L. G., Skirboll, L. and Emson, P. (1982). Substance P-containing primary sensory neurons projecting to the inferior mesenteric ganglion: evidence from combined retrograde tracing and immunohistochemistry. Neuroscience, 7, 647–654.CrossRefGoogle ScholarPubMed
Daly, M.d. (1991). Some reflex cardioinhibitory responses in the cat and their modulation by central inspiratory neuronal activity. J. Physiol., 439, 559–577.CrossRefGoogle ScholarPubMed
Daly, M.d. and Robinson, B. H. (1968). An analysis of the reflex systemic vasodilator response elicited by lung inflation in the dog. J. Physiol., 195, 387–406.CrossRefGoogle ScholarPubMed
Daly, M.d., Hazzledine, J. L. and Ungar, A. (1967). The reflex effects of alterations in lung volume on systemic vascular resistance in the dog. J. Physiol., 188, 331–351.CrossRefGoogle Scholar
Daly, M.d., Ward, J. and Wood, L. M. (1987). The peripheral chemoreceptors and cardiovascular-respiratory integration. In The Neurobiology of the Cardiorespiratory System, ed. Taylor, E. W.. Manchester: Manchester University Press, pp. 342–368.Google Scholar
Damasio, A. R. (1994). Descartes' Error: Emotion, Reason and the Human Brain. New York: Avon Books.Google Scholar
Damasio, A. R. (1999). The Feeling of What Happens: Body and Emotion in the Making of Consciousness. Harcourt Brace: New York.Google Scholar
Damasio, A. R., Adolphs, R. and Damasio, H. (2003). The contributions of the lesion method to the functional neuroanatomy of emotion. In Handbook of Affective Sciences, eds. Davidson, R. J., Scherer, K. R. and Goldsmith, H. H.. Oxford, New York: Oxford University Press, pp. 66–92.Google Scholar
Dampney, R. A. (1981). Brain stem mechanisms in the control of arterial pressure. Clin. Exp. Hypertens., 3, 379–391.CrossRefGoogle ScholarPubMed
Dampney, R. A. (1994). Functional organization of central pathways regulating the cardiovascular system. Physiol. Rev., 74, 323–364.CrossRefGoogle ScholarPubMed
Dampney, R. A. and Horiuchi, J. (2003). Functional organisation of central cardiovascular pathways: studies using c-fos gene expression. Prog. Neurobiol., 71, 359–384.CrossRefGoogle ScholarPubMed
Dampney, R. A. and McAllen, R. M. (1988). Differential control of sympathetic fibres supplying hindlimb skin and muscle by subretrofacial neurones in the cat. J. Physiol., 395, 41–56.CrossRefGoogle ScholarPubMed
Dampney, R. A. and Moon, E. A. (1980). Role of ventrolateral medulla in vasomotor response to cerebral ischemia. Am. J. Physiol., 239, H349–H358.Google ScholarPubMed
Dampney, R. A., Goodchild, A. K., Robertson, L. G. and Montgomery, W. (1982). Role of ventrolateral medulla in vasomotor regulation: a correlative anatomical and physiological study. Brain Res., 249, 223–235.CrossRefGoogle ScholarPubMed
Dampney, R. A., Goodchild, A. K. and Tan, E. (1985). Vasopressor neurons in the rostral ventrolateral medulla of the rabbit. J. Auton. Nerv. Syst., 14, 239–254.CrossRefGoogle ScholarPubMed
Dampney, R. A., Tagawa, T., Horiuchi, J.et al. (2000). What drives the tonic activity of presympathetic neurons in the rostral ventrolateral medulla?Clin. Exp. Pharmacol. Physiol., 27, 1049–1053.CrossRefGoogle ScholarPubMed
Dampney, R. A., Horiuchi, J., Tagawa, T.et al. (2003). Medullary and supramedullary mechanisms regulating sympathetic vasomotor tone. Acta Physiol. Scand., 177, 209–218.CrossRefGoogle ScholarPubMed
Dantzer, R., Bluthe, R. M., Gheusi, G.et al. (1998). Molecular basis of sickness behavior. Ann. New York Acad. Sci., 856, 132–138.CrossRefGoogle ScholarPubMed
Dantzer, R., Konsman, J. P., Bluthe, R. M. and Kelley, K. W. (2000). Neural and humoral pathways of communication from the immune system to the brain: parallel or convergent?Auton. Neurosci., 85, 60–65.CrossRefGoogle ScholarPubMed
Darwin, C. (1998) [1872]. The Expression of the Emotions in Man and Animals. With an Introduction, Afterword and Commentaries by Paul Ekman, 3rd edn. London: Harper Collins.Google Scholar
Davidson, R. J., Scherer, K. R. and Goldsmith, H. H. (eds.) (2003). Handbook of Affective Sciences. New York, Oxford: Oxford University Press.Google Scholar
Davies, P. J., Ireland, D. R. and McLachlan, E. M. (1996). Sources of Ca2+ for different Ca2 + -activated K+ conductances in neurones of the rat superior cervical ganglion. J. Physiol., 495, 353–366.CrossRefGoogle ScholarPubMed
Davies, P. J., Ireland, D. R., Martinez-Pinna, J. and McLachlan, E. M. (1999). Electrophysiological roles of L-type channels in different classes of guinea pig sympathetic neuron. J. Neurophysiol., 82, 818–828.CrossRefGoogle ScholarPubMed
Davis, K. D., Meyer, R. A. and Campbell, J. N. (1993). Chemosensitivity and sensitization of nociceptive afferents that innervate the hairy skin of monkey. J. Neurophysiol., 69, 1071–1081.CrossRefGoogle ScholarPubMed
Davis, M. J. and Hill, M. A. (1999). Signaling mechanisms underlying the vascular myogenic response. Physiol. Rev., 79, 387–423.CrossRefGoogle ScholarPubMed
Groat, W. C. (1976). Mechanisms underlying recurrent inhibition in the sacral parasympathetic outflow to the urinary bladder. J. Physiol., 257, 503–513.CrossRefGoogle ScholarPubMed
Groat, W. C. (1987). Neuropeptides in pelvic afferent pathways. Experientia, 43, 801–813.CrossRefGoogle ScholarPubMed
de Groat, W. C. (1989). Neuropeptides in pelvic afferent pathways. In Regulatory Peptides, ed. Polak, J. M.. Basel: Birkhauser Verlag AG, pp. 334–361.CrossRefGoogle Scholar
de Groat, W. C. (2002). Neural control of the urinary bladder and sexual organs. In Autonomic Failure, 4th edn., eds. Mathias, C. J. and Bannister, R.. New York, Oxford: Oxford University Press, pp. 151–165.Google ScholarPubMed
de Groat, W. C. and Booth, A. M. (1993). Neural control of penile erection. In The Autonomic Nervous System, Vol. 3, Nervous Control of the Urogenital System, ed. Maggi, C. A.. Chur, Switzerland: Harwood Academic Publishers, pp. 467–524.Google Scholar
Groat, W. C. and Ryall, R. W. (1968). Recurrent inhibition in sacral parasympathetic pathways to the bladder. J. Physiol., 196, 579–591.CrossRefGoogle Scholar
Groat, W. C., Booth, A. M., Milne, R. J. and Roppolo, J. R. (1982). Parasympathetic preganglionic neurons in the sacral spinal cord. J. Auton. Nerv. Syst., 5, 23–43.CrossRefGoogle ScholarPubMed
de Groat, W. C., Booth, A. M. and Yoshimura, N. (1993). Neurophysiology of micturition and its modification in animal models in human disease. In The Autonomic Nervous System, Vol. 3, Nervous Control of the Urogenital System. ed. Maggi, C. A.. Chur, Switzerland: Harwood Academic Publishers, pp. 227–290.Google Scholar
Groat, W. C., Vizzard, M. A., Araki, I. and Roppolo, J. H. (1996). Spinal interneurons and preganglionic neurons in sacral autonomic reflex pathways. Prog. Brain Res., 107, 97–111.CrossRefGoogle Scholar
Groat, W. C., Araki, I., Vizzard, M. A.et al. (1998). Developmental and injury induced plasticity in the micturition reflex pathway. Behav. Brain Res., 92, 127–140.CrossRefGoogle ScholarPubMed
Groat, W. C., Fraser, M. O., Yoshiyama, M.et al. (2001). Neural control of the urethra. Scand. J. Urol. Nephrol., Suppl (207), 35–43.Google ScholarPubMed
Delius, W., Hagbarth, K. E., Hongell, A. and Wallin, B. G. (1972). General characteristics of sympathetic activity in human muscle nerves. Acta Physiol. Scand., 84, 65–81.CrossRefGoogle ScholarPubMed
Dembowsky, K., Czachurski, J. and Seller, H. (1985). Morphology of sympathetic preganglionic neurons in the thoracic spinal cord of the cat: an intracellular horseradish peroxidase study. J. Comp. Neurol., 238, 453–465.CrossRefGoogle Scholar
Demir, S. S., Clark, J. W. and Giles, W. R. (1999). Parasympathetic modulation of sinoatrial node pacemaker activity in rabbit heart: a unifying model. Am. J. Physiol., 276, H2221–H2244.Google ScholarPubMed
Denton, K. M., Luff, S. E., Shweta, A. and Anderson, W. P. (2004). Differential neural control of glomerular ultrafiltration. Clin. Exp. Pharmacol. Physiol., 31, 380–386.CrossRefGoogle ScholarPubMed
Deuchars, S. A., Spyer, K. M. and Gilbey, M. P. (1997). Stimulation within the rostral ventrolateral medulla can evoke monosynaptic γ-aminobutyric acidergic inhibitory postsynaptic potentials in sympathetic preganglionic neurons in vitro. J. Neurophysiol., 77, 229–235.CrossRefGoogle Scholar
Deuchars, J., Li, Y. W., Kasparov, S. and Paton, J. F. (2000). Morphological and electrophysiological properties of neurones in the dorsal vagal complex of the rat activated by arterial baroreceptors. J. Comp. Neurol., 417, 233–249.3.0.CO;2-V>CrossRefGoogle ScholarPubMed
Deuchars, S. A., Brooke, R. E. and Deuchars, J. (2001a). Adenosine A1 receptors reduce release from excitatory but not inhibitory synaptic inputs onto lateral neurons. J. Neurosci., 21, 6308–6320.CrossRefGoogle Scholar
Deuchars, S. A., Brooke, R. E., Frater, B. and Deuchars, J. (2001b). Properties of interneurones in the intermediolateral cell column of the rat spinal cord: role of the potassium channel subunit Kv3.1. Neuroscience, 106, 433–446.CrossRefGoogle Scholar
Deuchars, S. A., Milligan, C. J., Stornetta, R. L. and Deuchars, J. (2005). γ-aminobutyric acidergic neurons in the central region of the spinal cord: a novel substrate for sympathetic inhibition. J. Neurosci., 25, 1063–1070.CrossRefGoogle Scholar
DiBona, G. F. and Kopp, U. C. (1997). Neural control of renal function. Physiol. Rev., 77, 75–197.CrossRefGoogle ScholarPubMed
Dickens, E. J., Hirst, G. D. and Tomita, T. (1999). Identification of rhythmically active cells in guinea-pig stomach. J. Physiol., 514, 515–531.CrossRefGoogle ScholarPubMed
Dietz, N. M., Rivera, J. M., Eggener, S. E.et al. (1994). Nitric oxide contributes to the rise in forearm blood flow during mental stress in humans. J. Physiol., 480, 361–368.CrossRefGoogle ScholarPubMed
Dittmar, C. (1873). Über die Lage des sogenannten Gefässcentrums in der Medulla oblongata [On the location of the so-called vascular center in the medulla oblongata]. Sitzungsber. Akad. Wiss. Wien, Math.-Naturw., Abt. 2, 25, 449–469.Google Scholar
Dobbins, E. G. and Feldman, J. L. (1994). Brainstem network controlling descending drive to phrenic motoneurons in rat. J. Comp. Neurol., 347, 64–86.CrossRefGoogle ScholarPubMed
Dockray, G. J., Green, T. and Varro, A. (1989). The afferent peptidergic innervation of the upper gastrointestinal tract. In Nerves and the Gastrointestinal Tract, eds. Singer, M. V. and Goebel, H.. Dordrecht: Kluwer Academic Publishers, pp. 105–122.Google Scholar
Dodt, C., Gunnarsson, T., Elam, M., Karlsson, T. and Wallin, B. G. (1995). Central blood volume influences sympathetic sudomotor nerve traffic in warm humans. Acta Physiol. Scand., 155, 41–51.CrossRefGoogle ScholarPubMed
Dodt, C., Lönnroth, P., Fehm, H. L. and Elam, M. (1999). Intraneural stimulation elicits an increase in subcutaneous interstitial glycerol levels in humans. J. Physiol., 521, 545–552.CrossRefGoogle ScholarPubMed
Donnerer, J., Schuligoi, R. and Stein, C. (1992). Increased content and transport of substance P and calcitonin gene-related peptide in sensory nerves innervating inflamed tissue: evidence for a regulatory function of nerve growth factor in vivo. Neuroscience, 49, 693–698.CrossRefGoogle ScholarPubMed
Dorward, P. K., Andresen, M. C., Burke, S. L., Oliver, J. R. and Korner, P. I. (1982). Rapid resetting of the aortic baroreceptors in the rabbit and its implications for short-term and longer term reflex control. Circ. Res., 50, 428–439.CrossRefGoogle ScholarPubMed
Dorward, P. K., Burke, S. L., Jänig, W. and Cassell, J. (1987). Reflex responses to baroreceptor, chemoreceptor and nociceptor inputs in single renal sympathetic neurones in the rabbit and the effects of anaesthesia on them. J. Auton. Nerv. Syst., 18, 39–54.CrossRefGoogle Scholar
Dostrovsky, J. O. and Craig, A. D. (2005). Ascending projection systems. In Wall and Melzack's Textbook of Pain. 5th edn., ed. McMahon, S. B. and Koltzenburg, M.. Edinburgh: Elsevier Churchill Livingstone, pp. 187–204.Google Scholar
Downing, J. E. and Miyan, J. A. (2000). Neural immunoregulation: emerging roles for nerves in immune homeostasis and disease. Immunol. Today, 21, 281–289.CrossRefGoogle ScholarPubMed
Drummond, P. D. (1995). Mechanisms of physiological gustatory sweating and flushing in the face. J. Auton. Nerv. Syst., 52, 117–124.CrossRefGoogle Scholar
Duan, Y. F., Winters, R., McCabe, P. M.et al. (1996). Behavioral characteristics of defense and vigilance reactions elicited by electrical stimulation of the hypothalamus in rabbits. Behav. Brain Res., 81, 33–41.CrossRefGoogle ScholarPubMed
Duan, Y. F., Winters, R., McCabe, P. M.et al. (1997). Functional relationship between the hypothalamic vigilance area and periaqueductal grey vigilance area. Physiol. Behav., 62, 675–679.CrossRefGoogle ScholarPubMed
Dun, N. J. (1983). Peptide hormones and transmission in sympathetic ganglia. In Autonomic Ganglia, ed. Elfvin, L. G.. Chichester: J. Wiley & Sons, pp. 345–366.Google Scholar
Dunn, W. R., Brock, J. A. and Hardy, T. A. (1999). Electrochemical and electrophysiological characterization of neurotransmitter release from sympathetic nerves supplying rat mesenteric arteries. Br. J. Pharmacol., 128, 174–180.CrossRefGoogle ScholarPubMed
Dworkin, B. R. (1993). Learning and Physiological Regulation. Chicago: The University of Chicago Press.Google Scholar
Dworkin, B. R. (2000). Interoception. In Handbook of Psychophysiology, 2nd edn., ed. Cacioppo, J. T., Tassinary, L. G. and Berntson, G. G.. Cambridge: Cambridge University Press, pp. 482–506.Google Scholar
Ebbeson, S. O. E. (1968a). Quantitative studies of superior cervical sympathetic ganglia in a variety of primates including man. I. The ratio of preganglionic neurons. J. Morphol., 124, 117–132.CrossRefGoogle Scholar
Ebbeson, S. O. E. (1968b). Quantitative studies of superior cervical sympathetic ganglia in a variety of primates including man. II. Neuronal packing density. J. Morphol., 124, 181–186.CrossRefGoogle Scholar
Eckberg, D. L. (2003). The human respiratory gate. J. Physiol., 548, 339–352.Google ScholarPubMed
Eckberg, D. L. and Sleight, P. (1992). Human Baroreceptor Reflexes in Health and Disease. Oxford: Clarendon Press.Google Scholar
Eckberg, D. L., Nerhed, C. and Wallin, B. G. (1985). Respiratory modulation of muscle sympathetic and vagal cardiac outflow in man. J. Physiol., 365, 181–196.CrossRefGoogle ScholarPubMed
Eckberg, D. L., Rea, R. F., Andersson, O. K.et al. (1988). Baroreflex modulation of sympathetic activity and sympathetic neurotransmitters in humans. Acta Physiol. Scand., 133, 221–231.CrossRefGoogle ScholarPubMed
Ectors, L. (1941). Contribution á l'étude des réactions pilomotrices. Arch. Int. Physiol., 51, 443–455.Google Scholar
Edwards, F. R., Bramich, N. J. and Hirst, G. D. (1993). Analysis of the effects of vagal stimulation on the sinus venous of the toad. Philos. Trans. R. Soc. London B Biol. Sci., 341, 149–162.CrossRefGoogle ScholarPubMed
Edwards, F. R., Hirst, G. D., Klemm, M. F. and Steele, P. A. (1995). Different types of ganglion cell in the cardiac plexus of guinea-pigs. J. Physiol., 486, 453–471.CrossRefGoogle ScholarPubMed
Edwards, S. L., Anderson, C. R., Southwell, B. R. and McAllen, R. M. (1996). Distinct preganglionic neurons innervate noradrenaline and adrenaline cells in the cat adrenal medulla. Neuroscience, 70, 825–832.CrossRefGoogle ScholarPubMed
Ekman, P. (1992). Facial expression of emotion: new findings, new questions. Psychol. Sci., 3, 34–38.CrossRefGoogle Scholar
Ekman, P. and Davidson, R. J. (eds.) (1994). The Nature of Emotions. Oxford: Oxford University Press.Google Scholar
Ekman, P., Levenson, R. W. and Friesen, M. V. (1983). Autonomic nervous system activity distinguishes between emotions. Science, 221, 1208–1210.CrossRefGoogle ScholarPubMed
Elfvin, L. G. (ed.) (1983). Autonomic Ganglia. Chichester: J. Wiley & Sons.Google Scholar
Elfvin, L. G., Lindh, B. and Hökfelt, T. (1993). The chemical neuroanatomy of sympathetic ganglia. Annu. Rev. Neurosci., 16, 471–507.CrossRefGoogle ScholarPubMed
Eliasson, S., Folkow, B., Lindgren, P. and Uvnäs, B. (1951). Activation of sympathetic vasodilator nerves to the skeletal muscles in the cat by hypothalamic stimulation. Acta Physiol. Scand., 23, 333–351.CrossRefGoogle ScholarPubMed
Ellison, G. D. and Zanchetti, A. (1973). Diffuse and specific activation of sympathetic cholinergic fibers of the cat. Am. J. Physiol., 225, 142–149.Google ScholarPubMed
Elsner, R., Franklin, D. L., Citters, R. L. and Kenney, D. W. (1966). Cardiovascular defense against asphyxia. Science, 153, 941–949.CrossRefGoogle ScholarPubMed
Emch, G. S., Hermann, G. E. and Rogers, R. C. (2000). tumor necrosis factor-alpha activates solitary nucleus neurons responsive to gastric distension. Am. J. Physiol. Gastrointest. Liver Physiol., 279, G582–G586.CrossRefGoogle ScholarPubMed
Emch, G. S., Hermann, G. E. and Rogers, R. C. (2002). Tumor necrosis factor-alpha inhibits physiologically identified dorsal motor nucleus neurons in vivo. Brain Res., 951, 311–315.CrossRefGoogle ScholarPubMed
Enquist, L. W. and Card, J. P. (2003). Recent advances in the use of neurotropic viruses for circuit analysis. Curr. Opin. Neurobiol., 13, 603–606.CrossRefGoogle ScholarPubMed
Eppel, G. A., Malpas, S. C., Denton, K. M. and Evans, R. G. (2004). Neural control of renal medullary perfusion. Clin. Exp. Pharmacol. Physiol., 31, 387–396.CrossRefGoogle ScholarPubMed
Ernsberger, U. (2000). Evidence for an evolutionary conserved role of bone morphogenetic protein growth factors and phox2 transcription factors during noradrenergic differentiation of sympathetic neurons. Induction of a putative synexpression group of neurotransmitter-synthesizing enzymes. Eur. J. Biochem., 267, 6976–6981.CrossRefGoogle ScholarPubMed
Ernsberger, U. (2001). The development of postganglionic sympathetic neurons: coordinating neuronal differentiation and diversification. Auton. Neurosci., 94, 1–13.CrossRefGoogle ScholarPubMed
Ernsberger, U., Esposito, L., Partimo, S.et al. (2005). Expression of neuronal markers suggests heterogeneity of chick sympathoadrenal cells prior to invasion of the adrenal anlagen. Cell Tissue Res., 319, 1–13.CrossRefGoogle ScholarPubMed
Esler, M., Jennings, G., Lambert, G.et al. (1990). Overflow of catecholamine neurotransmitters to the circulation: source, fate, and functions. Physiol. Rev., 70, 963–985.CrossRefGoogle ScholarPubMed
Evans, R. J. and Cunnane, T. C. (1992). Relative contributions of adenosine triphosphate and noradrenaline to the nerve evoked contraction of the rabbit jejunal artery. Dependence on stimulation parameters. Naunyn-Schmiedeberg's Arch. Pharmacol., 345, 424–430.CrossRefGoogle ScholarPubMed
Evans, R. J. and Surprenant, A. (1992). Vasoconstriction of guinea-pig submucosal arterioles following sympathetic nerve stimulation is mediated by the release of adenosine triphosphate. Br. J. Pharmacol., 106, 242–249.CrossRefGoogle Scholar
Ezure, K., Tanaka, I. and Kondo, M. (2003). Glycine is used as a transmitter by decrementing expiratory neurons of the ventrolateral medulla in the rat. J. Neurosci., 23, 8941–8948.CrossRefGoogle ScholarPubMed
Fagius, J. and Sundlöf, G. (1986). The diving response in man: effects on sympathetic activity in muscle and skin nerve fascicles. J. Physiol., 377, 429–443.CrossRefGoogle ScholarPubMed
Fagius, J. and Wallin, B. G. (1993). Long-term variability and reproducibility of resting human muscle nerve sympathetic activity at rest, as reassessed after a decade. Clin. Auton. Res., 3, 201–205.CrossRefGoogle Scholar
Fagius, J., Wallin, B. G., Sundlöf, G., Nerhed, C. and Englesson, S. (1985). Sympathetic outflow in man after anaesthesia of the glossopharyngeal and vagus nerves. Brain, 108, 423–438.CrossRefGoogle ScholarPubMed
Farkas, E., Jansen, A. S. and Loewy, A. D. (1998). Periaqueductal gray matter input to cardiac-related sympathetic premotor neurons. Brain Res., 792, 179–192.CrossRefGoogle ScholarPubMed
Feigl, E. O. (1998). Neural control of coronary blood flow. J. Vasc. Res., 35, 85–92.CrossRefGoogle ScholarPubMed
Feldberg, W. and Guertzenstein, P. G. (1976). Vasodepressor effects obtained by drugs acting on the ventral surface of the brain stem. J. Physiol., 258, 337–355.CrossRefGoogle ScholarPubMed
Feldman, J. L. (1986). Neurophysiology of breathing in mammals. In Handbook of Physiology, The Nervous System, Vol. IV, ed. Mountcastle, V. B. and Bloom, F. E., Bethesda: American Physiological Society, pp. 463–524.Google Scholar
Feldman, J. L. and McCrimmon, D. R. (2003). Neural control of breathing. In Fundamental Neuroscience, 2nd edn., ed. Squire, L. R., Bloom, F. E., McConnell, S. K.et al. San Diego: Academic Press, pp. 967–990.Google Scholar
Feldman, J. L., Mitchell, G. S. and Nattie, E. E. (2003). Breathing: rhythmicity, plasticity, chemosensitivity. Annu. Rev. Neurosci., 26, 239–266.CrossRefGoogle Scholar
Fernandez de Molina, A. and Hunsperger, R. W. (1962). Organization of the subcortical system governing defence and flight reactions in the cat. J. Physiol., 160, 200–213.CrossRefGoogle ScholarPubMed
Fielden, R., Sutton, T. J. and Taylor, E. M. (1980). Effect of tilting on the pressor responses to McN-A-343, a muscarinic sympathetic ganglion stimulant. Br. J. Pharmacol., 71, 287–295.CrossRefGoogle ScholarPubMed
Fields, H. L., Basbaum, A. I. and Heinricher, M. M. (2005). Central nervous system mechanisms of pain modulation. In Wall and Melzack's Textbook of Pain, 5th edn., eds. McMahon, S. B. and Koltzenburg, M.. Amsterdam, Edinburgh: Elsevier Churchill Livingstone, pp. 125–142.Google Scholar
Fleming, W. W. and Westfall, D. P. (1988). Adaptive supersensitivity. In Handbook of Experimental Pharmacology, Vol. 90/I, Catecholamines I, ed. Trendelenburg, U. and Weiner, N.. Berlin, Heidelberg, New York: Springer-Verlag, pp. 509–559.Google Scholar
Floyd, N. S., Price, J. L., Ferry, A. T., Keay, K. A. and Bandler, R. (2000). Orbitomedial prefrontal cortical projections to distinct longitudinal columns of the periaqueductal gray in the rat. J. Comp. Neurol., 422, 556–578.3.0.CO;2-U>CrossRefGoogle ScholarPubMed
Floyd, N. S., Price, J. L., Ferry, A. T., Keay, K. A. and Bandler, R. (2001). Orbitomedial prefrontal cortical projections to hypothalamus in the rat. J. Comp. Neurol., 432, 307–328.CrossRefGoogle ScholarPubMed
Foerster, O. (1927). Die Leitungsbahnen des Schmerzgefühls und die chirurgische Behandlung der Schmerzzustände [Pain Pathways and the Surgical Treatment of Pain]. Berlin, Wien: Urban und Schwarzenberg.Google Scholar
Fogel, R., Zhang, X. and Renehan, W. E. (1996). Relationships between the morphology and function of gastric and intestinal distention-sensitive neurons in the dorsal motor nucleus of the vagus. J. Comp. Neurol., 364, 78–91.3.0.CO;2-P>CrossRefGoogle ScholarPubMed
Folkow, B. (1955). The control of blood vessels. Physiol. Rev., 35, 629–663.CrossRefGoogle ScholarPubMed
Folkow, B. (1987). Psychosocial and central nervous influences in primary hypertension. Circulation, 76, I10–I19.Google ScholarPubMed
Folkow, B. (2000). Perspectives on the integrative function of the “sympatho-adrenomedullary system”. Auton. Neurosci., 83, 101–115.CrossRefGoogle Scholar
Folkow, B. and Euler, U. S. (1954). Selective activation of noradrenaline and adrenaline producing cells in the cat's adrenal gland by hypothalamic stimulation. Circ. Res., 2, 191–195.CrossRefGoogle ScholarPubMed
Folkow, B. and Neil, E. (1971). Circulation. New York: Oxford University Press.Google ScholarPubMed
Folkow, B. and Nilsson, H. (1997). Transmitter release at adrenergic nerve endings: Total exocytosis or fractional release?News Physiol. Sci., 12, 32–36.Google Scholar
Folkow, B., Schmidt, T. and Uvnäs-Moberg, K. (eds.) (1997). Stress, health and the social environment. Acta Physiol. Scand., 161, Suppl. 640, 1–179.Google Scholar
Forehand, C. J. (1987). Ultrastructural analysis of the distribution of synaptic boutons from labeled preganglionic axons on rabbit ciliary neurons. J. Neurosci., 7, 3274–3281.CrossRefGoogle ScholarPubMed
Forehand, C. J. and Purves, D. (1984). Regional innervation of rabbit ciliary ganglion cells by the terminals of preganglionic axons. J. Neurosci., 4, 1–12.CrossRefGoogle ScholarPubMed
Foreman, R. D. (1989). Organization of the spinothalamic tract as a relay for cardiopulmonary sympathetic afferent fiber activity. Prog. Sens. Physiol., 9, 1–51.CrossRefGoogle Scholar
Foreman, R. D. (1999). Mechanisms of cardiac pain. Annu. Rev. Physiol., 61, 143–167.CrossRefGoogle ScholarPubMed
Foster, M. (1879). Textbook of Physiology. In The Discovery of Reflexes, ed. Liddell, E. G. T.. Oxford: Clarendon Press, pp. 98–101.Google Scholar
Fox, E. A. and Powley, T. L. (1985). Longitudinal columnar organization within the dorsal motor nucleus represents separate branches of the abdominal vagus. Brain Res., 341, 269–282.CrossRefGoogle ScholarPubMed
Fox, E. A. and Powley, T. L. (1992). Morphology of identified preganglionic neurons in the dorsal motor nucleus of the vagus. J. Comp. Neurol., 322, 79–98.CrossRefGoogle ScholarPubMed
Frayn, K. N. and MacDonald, I. A. (1996). Adipose tissue circulation. In The Autonomic Nervous System, Vol. 8, Nervous Control of Blood Vessels, ed. Bennett, T. and Gardiner, S. M.. Amsterdam: Harwood Academic Publishers, pp. 505–539.Google Scholar
Freeman, R. and Rutkove, S. (2000). Syncope. In Handbook of Clinical Neurology, Vol. 75, The Autonomic Nervous System, part II: Dysfunctions, ed. Appenzeller, O.. Amsterdam: Elsevier, pp. 203–228.Google Scholar
Freyburger, W. A., Gruhzit, C. C. and Moe, G. K. (1950a). Pressor pathways not blocked by tetraethylammonium. Am. J. Physiol., 163, 290–293.Google Scholar
Freyburger, W. A., Gruhzit, C. C., Rennick, B. R. and Moe, G. K. (1950b). Action of tetraethylammonium on pressor response to asphyxia. Am. J. Physiol., 163, 554–560.Google Scholar
Fukai, K. and Fukuda, H. (1985). Three serial neurones in the innervation of the colon by the sacral parasympathetic nerve of the dog. J. Physiol., 362, 69–78.CrossRefGoogle ScholarPubMed
Fukami, H. and Bradley, R. M. (2005). Biophysical and morphological properties of parasympathetic neurons controlling the parotid and von Ebner salivary glands in rats. J. Neurophysiol., 93, 678–686.CrossRefGoogle ScholarPubMed
Fukuda, A., Minami, T., Nabekura, J. and Oomura, Y. (1987). The effects of noradrenaline on neurones in the rat dorsal motor nucleus of the vagus, in vitro. J. Physiol., 393, 213–231.CrossRefGoogle ScholarPubMed
Fulton, J. F. (1949). Physiology of the Nervous System, 3rd edn. New York: Oxford University Press.Google Scholar
Furness, J. B. (2005). The Enteric Nervous System. Oxford: Blackwell Science Ltd.Google Scholar
Furness, J. B. and Clerc, N. (2000). Responses of afferent neurons to the contents of the digestive tract, and their relation to endocrine and immune responses. Prog. Brain Res., 122, 159–172.CrossRefGoogle ScholarPubMed
Furness, J. B. and Costa, M. (1980). Types of nerves in the enteric nervous system. Neuroscience, 5, 1–20.Google Scholar
Furness, J. B. and Costa, M. (1987). The Enteric Nervous System. London: Churchill Livingstone.Google Scholar
Furness, J. B., Morris, J. L., Gibbins, I. L. and Costa, M. (1989). Chemical coding of neurons and plurichemical transmission. Annu. Rev. Pharmacol. Toxicol., 29, 289–306.CrossRefGoogle ScholarPubMed
Furness, J. B., Bornstein, J. C., Murphy, R. and Pompolo, S. (1992). Roles of peptides in transmission in the enteric nervous system. Trends Neurosci., 15, 66–71.CrossRefGoogle ScholarPubMed
Furness, J. B., Kunze, W. A., Bertrand, P. P., Clerc, N. and Bornstein, J. C. (1998). Intrinsic primary afferent neurons of the intestine. Prog. Neurobiol., 54, 1–18.CrossRefGoogle ScholarPubMed
Furness, J. B., Clerc, N., Vogalis, F. and Stebbing, M. J. (2003a). The enteric nervous system and its extrinsic connections. In Texbook of Gastroenterology, ed. Yamada, T., Alpers, D. H., Laine, L., Owyang, C. and Powell, D. W.. Philadelphia: Lippincott Williams & Wilkins, pp. 12–34.Google Scholar
Furness, J. B., Stebbing, M. J., Kunze, W. A. A. and Clerc, N. (2003b). Sensory neurons of the gastrointestinal tract. In Textbook of Gastroenterology, ed. Yamada, T., Alpers, D. H., Laine, L., Owyang, C. and Powell, D. W.. Philadelphia: Lippincott Williams & Wilkins, pp. 34–47.Google Scholar
Furness, J. B., Jones, C., Nurgali, K. and Clerc, N. (2004). Intrinsic primary afferent neurons and nerve circuits within the intestine. Prog. Neurobiol., 72, 143–164.CrossRefGoogle ScholarPubMed
Gabella, G. (1976). Structure of the Autonomic Nervous System. London: Chapman and Hall.CrossRefGoogle Scholar
Gamlin, P. D. R. (2000). Functions of the Edinger–Westphal nucleus. In The Autonomic Nervous System, Vol. 13, Nervous Control of the Eye, ed. Burnstock, G. and Sillito, A. M.. Amsterdam: Harwood Academic Publisher, pp. 117–154.Google Scholar
Gamlin, P. D. R. and Clarke, R. J. (1995). Single-unit activity in the primate nucleus reticularis tegmenti pontis related to vergence and ocular accommodation. J. Neurophysiol., 73, 2115–2119.CrossRefGoogle ScholarPubMed
Gamlin, P. D. R., Zhang, Y., Clendaniel, R. A. and Mays, L. E. (1994). Behavior of identified Edinger–Westphal neurons during ocular accommodation. J. Neurophysiol., 72, 2368–2382.CrossRefGoogle ScholarPubMed
Gandevia, S. C., Killian, K., McKenzie, D. K.et al. (1993). Respiratory sensations, cardiovascular control, kinaesthesia and transcranial stimulation during paralysis in humans. J. Physiol., 470, 85–107.CrossRefGoogle ScholarPubMed
Garrett, J. R., Ekström, J. and Anderson, L. C. (eds.) (1999). Frontiers in Oral Biology, Vol. 11, Basel: Karger.Google Scholar
Gaskell, W. H. (1916). The Involuntary Nervous System. London: Longmans.Google Scholar
Gauriau, C. and Bernard, J. F. (2002). Pain pathways and parabrachial circuits in the rat. Exp. Physiol., 87, 251–258.CrossRefGoogle ScholarPubMed
Gauriau, C. and Bernard, J. F. (2004a). Posterior triangular thalamic neurons convey nociceptive messages to the secondary somatosensory and insular cortices in the rat. J. Neurosci., 24, 752–761.CrossRefGoogle Scholar
Gauriau, C. and Bernard, J. F. (2004b). A comparative reappraisal of projections from the superficial laminae of the dorsal horn in the rat: the forebrain. J. Comp. Neurol., 468, 24–56.CrossRefGoogle Scholar
Gebhart, G. F. (ed.) (1995). Visceral Pain. Progress in Pain Research and Management. Seattle: IASP Press.Google Scholar
Gebhart, G. F. (1996). Visceral polymodal receptors. Prog. Brain Res., 113, 101–112.CrossRefGoogle ScholarPubMed
Gebhart, G. F. and Randich, A. (1992). Vagal modulation of nociception. Am. Pain Soc. J., 1, 26–32.Google Scholar
Gebber, G. L. (1990). Central determinants of sympathetic nerve discharge. In Central Regulation of Autonomic Functions, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 126–144.Google Scholar
Geerling, J. C., Mettenleiter, T. C. and Loewy, A. D. (2003). Orexin neurons project to diverse sympathetic outflow systems. Neuroscience, 122, 541–550.CrossRefGoogle ScholarPubMed
Gerber, U. and Polosa, C. (1978). Effects of pulmonary stretch receptor afferent stimulation on sympathetic preganglionic neuron firing. Can. J. Physiol. Pharmacol., 56, 191–198.CrossRefGoogle ScholarPubMed
Gerfen, C. R. and Sawchenko, P. E. (1984). An anterograde neuroanatomical tracing method that shows the detailed morphology of neurons, their axons and terminals: immunohistochemical localization of an axonally transported plant lectin, Phaseolus vulgaris leucoagglutinin (phaseolus vulgaris leuco-agglutinin). Brain Res., 290, 219–238.CrossRefGoogle Scholar
Gershon, M. D. (1994). Functional anatomy of the enteric nervous system. In Physiology of the Gastrointestinal Tract, 3rd edn., ed. Johnson, L. R.. New York: Raven Press, pp. 381–422.Google Scholar
Giamberardino, M. A. (1999). Recent and forgotten aspects of visceral pain. Eur. J. Pain, 3, 77–92.CrossRefGoogle ScholarPubMed
Gibbins, I. L. (1990). Target-related patterns of co-existence of neuropeptide Y, vasoactive intestinal peptide, enkephalin and substance P in cranial parasympathetic neurons innervating the facial skin and glands of guinea-pigs. Neuroscience, 38, 541–560.CrossRefGoogle ScholarPubMed
Gibbins, I. L. (1991). Vasomotor, pilomotor and secretomotor neurons distinguished by size and neuropeptide content in superior cervical ganglia of mice. J. Auton. Nerv. Syst., 34, 171–183.CrossRefGoogle Scholar
Gibbins, I. L. (1992). Vasoconstrictor, vasodilator and pilomotor pathways in sympathetic ganglia of guinea-pigs. Neuroscience, 47, 657–672.CrossRefGoogle ScholarPubMed
Gibbins, I. L. (1994). Comparative anatomy and evolution of the autonomic nervous system. In The Autonomic Nervous System, Vol. 4, Comparative Physiology and Evolution of the Autonomic Nervous System, ed. Nilsson, S. and Holmgren, S.. Chur: Harwood Academic Publishers, pp. 1–67.Google Scholar
Gibbins, I. L. (1995). Chemical neuroanatomy of sympathetic ganglia. In The Autonomic Nervous System, Vol. 6, Autonomic Ganglia, ed. McLachlan, E. M.. Luxembourg: Harwood Academic Publishers, pp. 73–122.Google Scholar
Gibbins, I. L. (1997). Autonomic pathways to cutaneous effectors. In The Autonomic Nervous System, Vol. 12, Autonomic Innervation of the Skin, ed. Morris, J. L. and Gibbins, I. L.. Chur, Switzerland: Harwood Academic Publishers, pp. 1–56.Google Scholar
Gibbins, I. L. (2004). Peripheral autonomic pathways. In The Human Nervous System, 2nd edn., ed. Paxinos, G. and Mai, J. K.. Amsterdam, San Diego, London: Elsevier Academic Press, pp. 134–189.Google Scholar
Gibbins, I. L. and Morris, J. L. (1987). Co-existence of neuropeptides in sympathetic, cranial autonomic and sensory neurons innervating the iris of the guinea-pig. J. Auton. Nerv. Syst., 21, 67–82.CrossRefGoogle ScholarPubMed
Gibbins, I. L. and Morris, J. L. (1990). Sympathetic noradrenergic neurons containing dynorphin but not neuropeptide Y innervate small cutaneous blood vessels of guinea-pigs. J. Auton. Nerv. Syst., 29, 137–149.CrossRefGoogle Scholar
Gibbins, I. L., Rodgers, H. F., Matthew, S. E. and Murphy, S. M. (1998). Synaptic organisation of lumbar sympathetic ganglia of guinea pigs: serial section ultrastructural analysis of dye-filled sympathetic final motor neurons. J. Comp. Neurol., 402, 285–302.3.0.CO;2-A>CrossRefGoogle ScholarPubMed
Gibbins, I. L., Jobling, P., Messenger, J. P., Teo, E. H. and Morris, J. L. (2000). Neuronal morphology and the synaptic organisation of sympathetic ganglia. J. Auton. Nerv. Syst., 81, 104–109.CrossRefGoogle ScholarPubMed
Gibbins, I. L., Jobling, P. and Morris, J. L. (2003a). Functional organization of peripheral vasomotor pathways. Acta Physiol. Scand., 177, 237–245.CrossRefGoogle Scholar
Gibbins, I. L., Jobling, P., Teo, E. H., Matthew, S. E. and Morris, J. L. (2003b). Heterogeneous expression of SNAP-25 and synaptic vesicle proteins by central and peripheral inputs to sympathetic neurons. J. Comp. Neurol., 459, 25–43.CrossRefGoogle Scholar
Gibbins, I. L., Teo, E. H., Jobling, P. and Morris, J. L. (2003c). Synaptic density, convergence, and dendritic complexity of prevertebral sympathetic neurons. J. Comp. Neurol., 455, 285–298.CrossRefGoogle Scholar
Gilbey, M. P., Jordan, D., Richter, D. W. and Spyer, K. M. (1984). Synaptic mechanisms involved in the inspiratory modulation of vagal cardio-inhibitory neurones in the cat. J. Physiol., 356, 65–78.CrossRefGoogle ScholarPubMed
Gilliat, R. W. (1948). Vaso-constriction in the finger after deep inspiration. J. Physiol., 107, 76–88.CrossRefGoogle Scholar
Gilliat, R. W., Guttmann, L. and Whitteridge, D. (1948). Inspiratory vasoconstriction in patients after spinal injuries. J. Physiol., 107, 67–75.CrossRefGoogle Scholar
Giuliano, F., Allard, J., Compagnie, S.et al. (2001). Vaginal physiological changes in a model of sexual arousal in anesthetized rats. Am. J. Physiol. Regul. Integr. Comp. Physiol., 281, R140–R149.CrossRefGoogle Scholar
Gladwell, S. J. and Coote, J. H. (1999). Inhibitory and indirect excitatory effects of dopamine on sympathetic preganglionic neurones in the neonatal rat spinal cord in vitro. Brain Res., 818, 397–407.CrossRefGoogle ScholarPubMed
Goehler, L. E., Gaykema, R. P., Hansen, M. K.et al. (2000). Vagal immune-to-brain communication: a visceral chemosensory pathway. Auton. Neurosci., 85, 49–59.CrossRefGoogle ScholarPubMed
Gola, M. and Niel, J. P. (1993). Electrical and integrative properties of rabbit sympathetic neurones re-evaluated by patch clamping non-dissociated cells. J. Physiol., 460, 327–349.CrossRefGoogle ScholarPubMed
Goldstein, D. S. (1995). Stress, Catecholamines, and Cardiovascular Disease. New York, Oxford: Oxford University Press.Google Scholar
Goldstein, D. S. (2000). The Autonomic Nervous System in Health and Disease. New York: Marcel Dekker.Google Scholar
Golenhofen, K., Hensel, H. and Ruef, J. (1962). Sustained dilatation in human muscle blood vessels under the influence of adrenaline. J. Physiol., 160, 189–199.CrossRefGoogle ScholarPubMed
Goodchild, A. K., Moon, E. A., Dampney, R. A. and Howe, P. R. (1984). Evidence that adrenaline neurons in the rostral ventrolateral medulla have a vasopressor function. Neurosci. Lett., 45, 267–272.CrossRefGoogle ScholarPubMed
Goodchild, A. K., Deurzen, B. T. M., Sun, Q. J., Chalmers, J. and Pilowsky, P. (2000). Spinal γ-aminobutyric acid receptors do not mediate the sympathetic baroreceptor reflex in the rat. Am. J. Physiol. Regul. Integr. Comp. Physiol., 279, R320–R331.CrossRefGoogle Scholar
Goodwin, G. M., McCloskey, D. I. and Mitchell, J. H. (1972). Cardiovascular and respiratory responses to changes in central command during isometric exercise at constant muscle tension. J. Physiol., 226, 173–190.CrossRefGoogle ScholarPubMed
Gordon, F. J. and McCann, L. A. (1988). Pressor responses evoked by microinjections of L-glutamate into the caudal ventrolateral medulla of the rat. Brain Res., 457, 251–258.CrossRefGoogle ScholarPubMed
Gore, A. C. and Roberts, J. L. (2003). Neuroendocrine systems. In Fundamental Neuroscience, 2nd edn., ed. Squire, L. R., Bloom, F. E., McConnell, S. K.et al. San Diego: Academic Press, pp. 1031–1065.Google Scholar
Goridis, C. and Rohrer, H. (2002). Specification of catecholaminergic and serotonergic neurons. Nat. Rev. Neurosci., 3, 531–541.CrossRefGoogle ScholarPubMed
Gould, D. J. and Hill, C. E. (1996). Alpha-adrenoceptor activation of a chloride conductance in rat iris arterioles. Am. J. Physiol., 271, H2469–H2476.Google ScholarPubMed
Granit, R. (1981). Comments on history of motor control. In Handbook of Physiology, Section I, The Nervous System, Vol. II, Motor Control, part I. ed. Brooks, V. B.. Bethesda: American Physiological Society, pp. 1–16.Google Scholar
Grasby, D. J., Gibbins, I. L. and Morris, J. L. (1997). Projections of sympathetic non-noradrenergic neurons to skeletal muscle arteries in guinea-pig limbs vary with the metabolic character of muscles. J. Vasc. Res., 34, 351–364.CrossRefGoogle ScholarPubMed
Gray's Anatomy (1995). 38th edn. Edinburgh: Churchill Livingstone.
Graziano, A. and Jones, E. G. (2004). Widespread thalamic terminations of fibers arising in the superficial medullary dorsal horn of monkeys and their relation to calbindin immunoreactivity. J. Neurosci., 24, 248–256.CrossRefGoogle ScholarPubMed
Green, P. G., Miao, F. J. P., Jänig, W. and Levine, J. D. (1995). Negative feedback neuroendocrine control of the inflammatory response in rats. J. Neurosci., 15, 4678–4686.CrossRefGoogle ScholarPubMed
Green, P. G., Jänig, W. and Levine, J. D. (1997). Negative feedback neuroendocrine control of inflammatory response in the rat is dependent on the sympathetic postganglionic neuron. J. Neurosci., 17, 3234–3238.CrossRefGoogle ScholarPubMed
Greenfield, A. D. (1966). Survey of the evidence for active neurogenic vasodilation in man. Fed. Proc., 25, 1607–1610.Google Scholar
Greger, R. and Windhorst, U. (eds.) (1996). Comprehensive Human Physiology. From Cellular Mechanisms to Integration, Vol. 1 and 2. Berlin, Heidelberg, New York: Springer Verlag.CrossRefGoogle Scholar
Gregor, M. and Jänig, W. (1977). Effects of systemic hypoxia and hypercapnia on cutaneous and muscle vasoconstrictor neurones to the cat's hindlimb. Pflügers Arch., 368, 71–81.CrossRefGoogle ScholarPubMed
Gregor, M., Jänig, W. and Riedel, W. (1976). Response pattern of cutaneous postganglionic neurones to the hindlimb on spinal cord heating and cooling in the cat. Pflügers Arch., 363, 135–140.CrossRefGoogle ScholarPubMed
Grewe, W., Jänig, W. and Kümmel, H. (1995). Effects of hypothalamic thermal stimuli on sympathetic neurones innervating skin and skeletal muscle of the cat hindlimb. J. Physiol., 488, 139–152.CrossRefGoogle ScholarPubMed
Griffith, W. H. III, Gallagher, J. P. and Shinnick-Gallagher, P. (1980). An intracellular investigation of cat vesical pelvic ganglia. J. Neurophysiol., 43, 343–354.CrossRefGoogle ScholarPubMed
Grigg, P., Schaible, H. G. and Schmidt, R. F. (1986). Mechanical sensitivity of group III and IV afferents from posterior articular nerve in normal and inflamed cat knee. J. Neurophysiol., 55, 635–643.CrossRefGoogle ScholarPubMed
Grosse, M. and Jänig, W. (1976). Vasoconstrictor and pilomotor fibres in skin nerves to the cat's tail. Pflügers Arch., 361, 221–229.CrossRefGoogle ScholarPubMed
Grubb, B. P. and Karas, B. (2002). Neurally mediated syncope. In Autonomic Failure, 4th edn., ed. Mathias, C. J. and Bannister, R.. Oxford: Oxford University Press, pp. 437–447.Google ScholarPubMed
Grundy, D. (1988). Speculation on the structure/function relationship for vagal and splanchnic afferent endings supplying the gastrointestinal tract. J. Auton. Nerv. Syst., 22, 175–180.CrossRefGoogle Scholar
Grundy, D. and Scratcherd, T. (1989). Sensory afferents from the gastrointestinal tract. In Handbook of Physiology, Section 6: The Gastrointestinal System, Vol. 1: Motility and Circulation, ed. Wood, J. D.. Bethesda: American Physiological Society, pp. 593–620.Google Scholar
Grundy, D., Salih, A. A. and Scratcherd, T. (1981). Modulation of vagal efferent fibre discharge by mechanoreceptors in the stomach, duodenum and colon of the ferret. J. Physiol., 319, 43–52.CrossRefGoogle ScholarPubMed
Guertzenstein, P. G. and Silver, A. (1974). Fall in blood pressure produced from discrete regions of the ventral surface of the medulla by glycine and lesions. J. Physiol., 242, 489–503.CrossRefGoogle ScholarPubMed
Guth, L. and Bernstein, J. J. (1961). Selectivity in the re-establishment of synapses in the superior cervical sympathetic ganglion of the cat. Exp. Neurol., 4, 59–69.CrossRefGoogle ScholarPubMed
Guttmann, L. (1976). Spinal Cord Injuries. Comprehensive Management and Research. Oxford: Blackwell Scientific Publications.Google Scholar
Guyenet, P. G. (1990). Role of the ventral medulla oblongata in blood pressure regulation. In Central Regulation of Autonomic Functions, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 145–167.Google Scholar
Guyenet, P. G. (2000). Neural structures that mediate sympathoexcitation during hypoxia. Respir. Physiol., 121, 147–162.CrossRefGoogle ScholarPubMed
Guyenet, P. G. and Brown, D. L. (1986). Nucleus paragigantocellularis lateralis and lumbar sympathetic discharge in the rat. Am. J. Physiol., 250, R1081–R1094.Google ScholarPubMed
Guyenet, P. G. and Koshiya, N. (1992). Respiratory-sympathetic integration in the medulla oblongata. In Central Neural Mechanisms in Cardiovascular Regulation, ed. Kunos, G. and Ciriello, J.. Boston: Birkhäuser, pp. 226–247.CrossRefGoogle Scholar
Guyenet, P. G. and Stornetta, R. L. (1997). Central nervous system regulation of the sympathetic and cardiovagal vasomotor outflows. In Anesthesia, ed. Yaksh, T. L., Maze, M., Lynch, C.et al. Philadelphia, New York: Lippincott-Raven, pp. 1205–1232.Google Scholar
Guyenet, P. G., Koshiya, N., Huangfu, D.et al. (1996). Role of medulla oblongata in generation of sympathetic and vagal outflows. Prog. Brain Res., 107, 127–144.CrossRefGoogle ScholarPubMed
Guyenet, P. G., Schreihofer, A. M. and Stornetta, R. L. (2001). Regulation of sympathetic tone and arterial pressure by the rostral ventrolateral medulla after depletion of C1 cells in rats. Ann. New York Acad. Sci., 940, 259–269.CrossRefGoogle ScholarPubMed
Häbler, H. J. and Jänig, W. (1995). Coordination of sympathetic and respiratory systems: neurophysiological experiments. Clin. Exp. Hypertens., 17, 223–235.CrossRefGoogle ScholarPubMed
Häbler, H. J., Jänig, W. and Koltzenburg, M. (1990a). Activation of unmyelinated afferent fibres by mechanical stimuli and inflammation of the urinary bladder in the cat. J. Physiol., 425, 545–562.CrossRefGoogle Scholar
Häbler, H. J., Jänig, W., Koltzenburg, M. and McMahon, S. B. (1990b). A quantitative study of the central projection patterns of unmyelinated ventral root afferents in the cat. J. Physiol., 422, 265–287.CrossRefGoogle Scholar
Häbler, H. J., Hilbers, K., Jänig, W.et al. (1992). Viscero-sympathetic reflex responses to mechanical stimulation of pelvic viscera in the cat. J. Auton. Nerv. Syst., 38, 147–158.CrossRefGoogle ScholarPubMed
Häbler, H. J., Jänig, W. and Koltzenburg, M. (1993a). Myelinated primary afferents of the sacral spinal cord responding to slow filling and distension of the urinary bladder. J. Physiol., 463, 449–460.CrossRefGoogle Scholar
Häbler, H. J., Jänig, W. and Koltzenburg, M. (1993b). Receptive properties of myelinated primary afferents innervating the inflamed urinary bladder of the cat. J. Neurophysiol., 69, 395–405.CrossRefGoogle Scholar
Häbler, H. J., Jänig, W., Krummel, M. and Peters, O. A. (1993c). Respiratory modulation of the activity in postganglionic neurons supplying skeletal muscle and skin of the rat hindlimb. J. Neurophysiol., 70, 920–930.CrossRefGoogle Scholar
Häbler, H. J., Jänig, W., Krummel, M. and Peters, O. A. (1994a). Reflex patterns in postganglionic neurons supplying skin and skeletal muscle of the rat hindlimb. J. Neurophysiol., 72, 2222–2236.CrossRefGoogle Scholar
Häbler, H. J., Jänig, W. and Michaelis, M. (1994b). Respiratory modulation of activity in sympathetic neurones. Prog. Neurobiol., 43, 567–606.CrossRefGoogle Scholar
Häbler, H. J., Bartsch, T. and Jänig, W. (1996). Two distinct mechanisms generate the respiratory modulation in fibre activity of the rat cervical sympathetic trunk. J. Auton. Nerv. Syst., 61, 116–122.CrossRefGoogle ScholarPubMed
Häbler, H. J., Boczek-Funcke, A., Michaelis, M. and Jänig, W. (1997a). Responses of distinct types of sympathetic neuron to stimulation of the superior laryngeal nerve in the cat. J. Auton. Nerv. Syst., 66, 97–104.CrossRefGoogle Scholar
Häbler, H. J., Wasner, G., Bartsch, T. and Jänig, W. (1997b). Responses of rat postganglionic sympathetic vasoconstrictor neurons following blockade of nitric oxide synthesis in vivo. Neuroscience, 77, 899–909.CrossRefGoogle Scholar
Häbler, H. J., Wasner, G. and Jänig, W. (1997c). Interaction of sympathetic vasoconstriction and antidromic vasodilatation in the control of skin blood flow. Exp. Brain Res., 113, 402–410.CrossRefGoogle Scholar
Häbler, H. J., Bartsch, T. and Jänig, W. (1999a). Rhythmicity in single fiber postganglionic activity supplying the rat tail. J. Neurophysiol., 81, 2026–2036.CrossRefGoogle Scholar
Häbler, H. J., Timmermann, L., Stegmann, J. U. and Jänig, W. (1999b). Involvement of neurokinins in antidromic vasodilatation in hairy and hairless skin of the rat hindlimb. Neuroscience, 89, 1259–1268.CrossRefGoogle Scholar
Häbler, H. J., Bartsch, T. and Jänig, W. (2000). Respiratory rhythmicity in the activity of postganglionic neurones supplying the rat tail during hyperthermia. J. Auton. Nerv. Syst., 83, 75–80.Google ScholarPubMed
Hadziefendic, S. and Haxhiu, M. A. (1999). central nervous system innervation of vagal preganglionic neurons controlling peripheral airways: a transneuronal labeling study using pseudorabies virus. J. Auton. Nerv. Syst., 76, 135–145.CrossRefGoogle ScholarPubMed
Hagbarth, K. E. and Vallbo, A. B. (1968). Pulse and respiratory grouping of sympathetic impulses in human muscle nerves. Acta Physiol. Scand., 74, 96–108.CrossRefGoogle ScholarPubMed
Hagbarth, K. E., Hallin, R. G., Hongell, A., Torebjörk, H. E. and Wallin, B. G. (1972). General characteristics of sympathetic activity in human skin nerves. Acta Physiol. Scand., 84, 164–176.CrossRefGoogle ScholarPubMed
Hainsworth, R. (2002). Syncope and fainting: classification pathophysiological basis. In Autonomic Failure, 4th edn., eds. Mathias, C. J. and Bannister, R.. Oxford: Oxford University Press, pp. 428–436.Google ScholarPubMed
Hall, M. (1841). On the Diseases and Derangements of the Nervous System, in their Primary Forms and their Modifications by Age, Sex Constitution, Heredity, Disposition, Excesses, General Disorder, and Organic Disease. London: Bailliere.Google Scholar
Halliday, G. M. and McLachlan, E. M. (1991). A comparative analysis of neurons containing catecholamine-synthesizing enzymes and neuropeptide Y in the ventrolateral medulla of rats, guinea-pigs and cats. Neuroscience, 43, 531–550.CrossRefGoogle Scholar
Hallin, R. G. and Torebjörk, H. E. (1974). Single unit sympathetic activity in human skin nerves during rest and various manoeuvres. Acta Physiol. Scand., 92, 303–317.CrossRefGoogle ScholarPubMed
Hamblin, P. A., McLachlan, E. M. and Lewis, R. J. (1995). Sub-nanomolar concentrations of ciguatoxin-1 excite preganglionic terminals in guinea pig sympathetic ganglia. Naunyn-Schmiedeberg's Arch. Pharmacol., 352, 236–246.CrossRefGoogle ScholarPubMed
Harden, R. N., Baron, R. and Jänig, W. (eds.) (2001). Complex Regional Pain Syndrome. Seattle: IASP Press.Google ScholarPubMed
Hargreaves, K. M., Roszkowski, M. T. and Swift, J. Q. (1993). Bradykinin and inflammatory pain. Agents Actions Suppl., 41, 65–73.Google ScholarPubMed
Harhun, M. I., Pucovsky, V., Povstyan, O. V., Gordienko, D. V. and Bolton, T. B. (2005). Interstitial cells in the vasculature. J. Cell Mol. Med., 9, 232–243.CrossRefGoogle ScholarPubMed
Haselton, J. R. and Guyenet, P. G. (1989). Central respiratory modulation of medullary sympathoexcitatory neurons in rat. Am. J. Physiol., 256, R739–R750.Google ScholarPubMed
Haupt, P., Jänig, W. and Kohler, W. (1983). Response pattern of visceral afferent fibres, supplying the colon, upon chemical and mechanical stimuli. Pflügers Arch., 398, 41–47.CrossRefGoogle ScholarPubMed
Haxhiu, M. A. and Loewy, A. D. (1996). Central connections of the motor and sensory vagal systems innervating the trachea. J. Auton. Nerv. Syst., 57, 49–56.CrossRefGoogle ScholarPubMed
Haxhiu, M. A., Jansen, A. S., Cherniack, N. S. and Loewy, A. D. (1993). central nervous system innervation of airway-related parasympathetic preganglionic neurons: a transneuronal labeling study using pseudorabies virus. Brain Res., 618, 115–134.CrossRefGoogle ScholarPubMed
Haxhiu, M. A., Erokwu, B., Bhardwaj, V. and Dreshaj, I. A. (1998). The role of the medullary raphe nuclei in regulation of cholinergic outflow to the airways. J. Auton. Nerv. Syst., 69, 64–71.CrossRefGoogle ScholarPubMed
Heel, K. A., McCauley, R. D., Papadimitriou, J. M. and Hall, J. C. (1997). Review: Peyer's patches. J. Gastroenterol. Hepatol., 12, 122–136.CrossRefGoogle ScholarPubMed
Hellmann, K. (1963). The effect of temperature changes on the isolated pilomotor muscles. J. Physiol., 169, 621–629.CrossRefGoogle ScholarPubMed
Henderson, C. G. and Ungar, A. (1978). Effect of cholinergic antagonists on sympathetic ganglionic transmission of vasomotor reflexes from the carotid baroreceptors and chemoreceptors of the dog. J. Physiol., 277, 379–385.CrossRefGoogle ScholarPubMed
Henderson, L. A., Keay, K. A. and Bandler, R. (1998). The ventrolateral periaqueductal gray projects to caudal brainstem depressor regions: a functional-anatomical and physiological study. Neuroscience, 82, 201–221.CrossRefGoogle ScholarPubMed
Hendry, I. A. and Hill, C. E. (eds.) (1992). The Autonomic Nervous System, Vol. 2, Development, Regeneration and Plasticity, Chur: Harwood Academic Publishers.Google Scholar
Henry, J. P. (1997). Culture and High Blood Pressure. Hamburg, Münster: LIT Publishing Company.Google Scholar
Henry, J. P. and Grim, C. E. (1990). Psychosocial mechanisms of primary hypertension. J. Hypertens., 8, 783–793.CrossRefGoogle ScholarPubMed
Henry, J. P. and Stephens, P. M. (1977). Stress, Health and the Social Environment: a Sociobiologic Approach to Medicine. Heidelberg, Berlin: Springer.CrossRefGoogle Scholar
Hensel, H. (1981). Thermoreception and Temperature Regulation. London, New York: Academic Press.Google ScholarPubMed
Hensel, H. (1982). Thermal Sensations and Thermoreceptors in Man. Springfield Illinois: Charles C. Thomas Publ.Google Scholar
Herbert, H., Moga, M. M. and Saper, C. B. (1990). Connections of the parabrachial nucleus with the nucleus of the solitary tract and the medullary reticular formation in the rat. J. Comp. Neurol., 293, 540–580.CrossRefGoogle ScholarPubMed
Hering, E. (1869). Über den Einfluβ der Atmung auf den Kreislauf. Erste Mitteilung. Über Atembewegungen der Gefäβ systeme [On the influence of the respiration on the circulation system. First report. On the respiratory movement of the vascular systems]. Sitzungsber. Akad. Wiss. Wien, Math.- Naturw., Abt. 2, 60, 829–856.Google Scholar
Hering, H. E. (1927). Die Karotissinusreflexe [The Carotid Sinus Reflexes]. Dresden, Leipzig: Verlag von Theodor Steinkopf.Google Scholar
Hermann, G. and Rogers, R. C. (1995). Tumor necrosis factor-alpha in the dorsal vagal complex suppresses gastric motility. NeuroImmunoModulation, 2, 74–81.CrossRefGoogle ScholarPubMed
Hermann, G. E., Tovar, C. A. and Rogers, R. C. (2002). LPS-induced suppression of gastric motility relieved by tumor necrosis factorR:Fc construct in dorsal vagal complex. Am. J. Physiol. Gastrointest. Liver Physiol., 283, G634–G639.CrossRefGoogle Scholar
Hertz, A. F. (1911). The Sensibility of the Alimentary Canal. Oxford: Oxford University Press.Google Scholar
Hess, W. R. (1944). Hypothalamische Adynamie [Hypothalamic adynamia]. Helv. Physiol. Acta, 2, 137–147.Google Scholar
Hess, W. R. (1948). Die Organisation des vegetativen Nervensystems [The Organization of the Autonomic Nervous System]. Basel: Benno Schwabe & Co.Google Scholar
Hess, W. R. (1954). Das Zwischenhirn, Syndrome, Lokalisationen, Funktionen [The Diencephalon, Syndromes, Localizations, Functions], 2nd edn. Basel: Benno Schwabe.Google Scholar
Hess, W. R. and Brügger, M. (1943). Das subcortikale Zentrum der affektiven Abwehrreaktion [The subcortical center of the affective defense reaction]. Helv. Physiol. Acta, 1, 33–52.Google Scholar
Heuckeroth, R. O., Enomoto, H., Grider, J. R.et al. (1999). Gene targeting reveals a critical role for neurturin in the development and maintenance of enteric, sensory, and parasympathetic neurons. Neuron, 22, 253–263.CrossRefGoogle ScholarPubMed
Heymans, C. and Neil, E. (1958). Reflexogenic Areas of the Cardiovascular System. London: J. & A. Churchill Ltd.Google Scholar
Hill, C. E., Hendry, I. A. and Sheppard, A. (1987). Use of the fluorescent dye, fast blue, to label sympathetic postganglionic neurones supplying mesenteric arteries and enteric neurones of the rat. J. Auton. Nerv. Syst., 18, 73–82.CrossRefGoogle ScholarPubMed
Hill, C. E., Klemm, M., Edwards, F. R. and Hirst, G. D. (1993). Sympathetic transmission to the dilator muscle of the rat iris. J. Auton. Nerv. Syst., 45, 107–123.CrossRefGoogle ScholarPubMed
Hill, C. E., Eade, J. and Sandow, S. L. (1999). Mechanisms underlying spontaneous rhythmical contractions in irideal arterioles of the rat. J. Physiol., 521, 507–516.CrossRefGoogle ScholarPubMed
Hill, M. A., Zou, H., Potocnik, S. J., Meininger, G. A. and Davis, M. J. (2001). Invited review: arteriolar smooth muscle mechanotransduction: Ca2 + signaling pathways underlying myogenic reactivity. J. Appl. Physiol., 91, 973–983.CrossRefGoogle Scholar
Hillsley, K. and Grundy, D. (1998). Sensitivity to 5-hydroxytryptamine in different afferent subpopulations within mesenteric nerves supplying the rat jejunum. J. Physiol., 509, 717–727.CrossRefGoogle ScholarPubMed
Hillsley, K., Kirkup, A. J. and Grundy, D. (1998). Direct and indirect actions of 5-hydroxytryptamine on the discharge of mesenteric afferent fibres innervating the rat jejunum. J. Physiol., 506, 551–561.CrossRefGoogle ScholarPubMed
Himms-Hagen, J. (1991). Neural control of brown adipose tissue: thermogenesis, hypertrophy, and atrophy neuroeffector junctions. Front. Neuroendocrinol., 12, 38–93.Google Scholar
Hirooka, Y., Polson, J. W., Potts, P. D. and Dampney, R. A. (1997). Hypoxia-induced Fos expression in neurons projecting to the pressor region in the rostral ventrolateral medulla. Neuroscience, 80, 1209–1224.CrossRefGoogle ScholarPubMed
Hirshberg, R. M., Al Chaer, E. D., Lawand, N. B., Westlund, K. N. and Willis, W. D. (1996). Is there a pathway in the posterior funiculus that signals visceral pain?Pain, 67, 291–305.CrossRefGoogle Scholar
Hirst, G. D. (2001). An additional role for interstitial cell of Cajal in the control of gastrointestinal motility?J. Physiol., 537, 1CrossRefGoogle Scholar
Hirst, G. D. and Edwards, F. R. (1989). Sympathetic neuroeffector transmission in arteries and arterioles. Physiol. Rev., 69, 546–604.CrossRefGoogle ScholarPubMed
Hirst, G. D. and McLachlan, E. M. (1984). Post-natal development of ganglia in the lower lumbar sympathetic chain of the rat. J. Physiol., 349, 119–134.CrossRefGoogle ScholarPubMed
Hirst, G. D. and McLachlan, E. M. (1986). Development of dendritic calcium currents in ganglion cells of the rat lower lumbar sympathetic chain. J. Physiol., 377, 349–368.CrossRefGoogle ScholarPubMed
Hirst, G. D. and Neild, T. O. (1980). Some properties of spontaneous excitatory junction potentials recorded from arterioles of guinea-pigs. J. Physiol., 303, 43–60.CrossRefGoogle ScholarPubMed
Hirst, G. D. and Ward, S. M. (2003). Interstitial cells: involvement in rhythmicity and neural control of gut smooth muscle. J. Physiol., 550, 337–346.CrossRefGoogle ScholarPubMed
Hirst, G. D. S., Holman, M. E. and McKirdy, H. C. (1974). Two types of neurones in the myenteric plexus of duodenum in the guinea-pig. J. Physiol., 236, 303–326.CrossRefGoogle ScholarPubMed
Hirst, G. D., Holman, M. E. and McKirdy, H. C. (1975). Two descending nerve pathways activated by distension of guinea-pig small intestine. J. Physiol., 244, 113–127.CrossRefGoogle ScholarPubMed
Hirst, G. D. S., Edwards, F. R., Bramich, N. J. and Klemm, M. F. (1991). Neural control of cardiac pacemaker potentials. News Physiol. Sci., 6, 185–190.Google Scholar
Hirst, G. D. S., Bramich, N. J., Edwards, F. R. and Klemm, M. (1992). Transmission at autonomic neuroeffector junctions. Trends Neurosci., 15, 40–46.Google ScholarPubMed
Hirst, G. D., Choate, J. K., Cousins, H. M., Edwards, F. R. and Klemm, M. F. (1996). Transmission by post-ganglionic axons of the autonomic nervous system: the importance of the specialized neuroeffector junction. Neuroscience, 73, 7–23.CrossRefGoogle ScholarPubMed
Hoffmeister, B., Hussels, W. and Jänig, W. (1978). Long-lasting discharge of postganglionic neurones to skin and muscle of the cat's hindlimb after repetitive activation of preganglionic axons in the lumbar sympathetic trunk. Pflügers Arch., 376, 15–20.CrossRefGoogle ScholarPubMed
Hohmann, E. L., Elde, R. P., Rysavy, J. A., Einzig, S. and Gebhard, R. L. (1986). Innervation of periosteum and bone by sympathetic vasoactive intestinal peptide-containing nerve fibers. Science, 232, 868–871.CrossRefGoogle ScholarPubMed
Holman, M. E., Coleman, H. A., Tonta, M. A. and Parkington, H. C. (1994). Synaptic transmission from splanchnic nerves to the adrenal medulla of guinea-pigs. J. Physiol., 478, 115–124.CrossRefGoogle ScholarPubMed
Holzer, P. (1992). Peptidergic sensory neurons in the control of vascular functions: mechanisms and significance in the cutaneous and splanchnic vascular beds. Rev. Physiol. Biochem. Pharmacol., 121, 49–146.CrossRefGoogle ScholarPubMed
Holzer, P. (1995). Chemosensitive afferent nerves in the regulation of gastric blood flow and protection. Adv. Exp. Med. Biol., 371B, 891–895.Google ScholarPubMed
Holzer, P. (1998a). Neural emergency system in the stomach. Gastroenterology, 114, 823–839.CrossRefGoogle Scholar
Holzer, P. (1998b). Neurogenic vasodilatation and plasma leakage in the skin. Gen. Pharmacol., 30, 5–11.CrossRefGoogle Scholar
Holzer, P. (2002a). Sensory neurone responses to mucosal noxae in the upper gut: relevance to mucosal integrity and gastrointestinal pain. Neurogastroenterol. Motil., 14, 459–475.CrossRefGoogle Scholar
Holzer, P. (2002b). Control of gastric functions by extrinsic sensory neurons. In The Autonomic Nervous System, Vol. 14, Innervation of the Gastrointestinal Tract, ed. Brookes, S. and Costa, M., London, New York: Taylor and Francis, pp. 103–170.Google Scholar
Holzer, P. (2003). Afferent signalling of gastric acid challenge. J. Physiol. Pharmacol., 54 Suppl4, 43–53.Google ScholarPubMed
Holzer, P. and Maggi, C. A. (1998). Dissociation of dorsal root ganglion neurons into afferent and efferent-like neurons. Neuroscience, 86, 389–398.Google ScholarPubMed
Hopkins, D. A., Bieger, D., deVente, J. and Steinbusch, W. M. (1996). Vagal efferent projections: viscerotopy, neurochemistry and effects of vagotomy. Prog. Brain Res., 107, 79–96.CrossRefGoogle ScholarPubMed
Horeyseck, G. and Jänig, W. (1974a). Reflexes in postganglionic fibres within skin and muscle nerves after mechanical non-noxious stimulation of skin. Exp. Brain Res., 20, 115–123.Google Scholar
Horeyseck, G. and Jänig, W. (1974b). Reflexes in postganglionic fibres within skin and muscle nerves after noxious stimulation of skin. Exp. Brain Res., 20, 125–134.Google Scholar
Horeyseck, G. and Jänig, W. (1974c). Reflex activity in postganglionic fibres within skin and muscle nerves elicited by somatic stimuli in chronic spinal cats. Exp. Brain Res., 21, 155–168.CrossRefGoogle Scholar
Horeyseck, G., Jänig, W., Kirchner, F. and Thämer, V. (1976). Activation and inhibition of muscle and cutaneous postganglionic neurones to hindlimb during hypothalamically induced vasoconstriction and atropine-sensitive vasodilation. Pflügers Arch., 361, 231–240.CrossRefGoogle ScholarPubMed
Hori, T., Katafuchi, T., Take, S., Shimizu, N. and Niijima, A. (1995). The autonomic nervous system as a communication channel between the brain and the immune system. NeuroImmunoModulation, 2, 203–215.CrossRefGoogle ScholarPubMed
Horiuchi, J. and Dampney, R. A. L. (2002). Evidence for tonic disinhibition of rostral ventrolateral medulla sympathoexcitatory neurons from the caudal pressor area. Auton. Neurosci., 99, 102–110.CrossRefGoogle ScholarPubMed
Horiuchi, J., Killinger, S. and Dampney, R. A. (2004a). Contribution to sympathetic vasomotor tone of tonic glutamatergic inputs to neurons in the rostral ventrolateral medulla. Am. J. Physiol. Regul. Integr. Comp. Physiol., 287, R1335–R1343.CrossRefGoogle Scholar
Horiuchi, J., McAllen, R. M., Allen, A. M.et al. (2004b). Descending vasomotor pathways from the dorsomedial hypothalamic nucleus: role of medullary raphe and rostral ventrolateral medulla. Am. J. Physiol. Regul. Integr. Comp. Physiol., 287, R824–R832.CrossRefGoogle Scholar
Horowitz, B., Ward, S. M. and Sanders, K. M. (1999). Cellular and molecular basis for electrical rhythmicity in gastrointestinal muscles. Annu. Rev. Physiol., 61, 19–43.CrossRefGoogle ScholarPubMed
Hosoya, Y., Sugiura, Y., Okado, N., Loewy, A. D. and Kohno, K. (1991). Descending input from the hypothalamic paraventricular nucleus to sympathetic preganglionic neurons in the rat. Exp. Brain Res., 85, 10–20.CrossRefGoogle ScholarPubMed
Hoyle, C. H. V., Milner, P. and Burnstock, G. (2002). Neuroeffector transmission in the intestine. In The Autonomic Nervous System, Vol. 14, Innervation of the Gastrointestinal Tract, ed. Brookes, S. and Costa, M.. London, New York: Taylor and Francis, pp. 295–340.Google Scholar
Hua, L. H., Strigo, I. A., Baxter, L. C., Johnson, S. C. and Craig, A. D. (2005). Anterior-posterior somatotopy of innocuous cooling activation focus in human dorsal insular cortex. Am. J. Physiol. Regul. Intrgr. Comp. Physiol., 289, R319–R325.CrossRefGoogle Scholar
Huang, J. and Weiss, M. L. (1999). Characterization of the central cell groups regulating the kidney in the rat. Brain Res., 845, 77–91.CrossRefGoogle ScholarPubMed
Huang, X. F., Törk, I. and Paxinos, G. (1993). Dorsal motor nucleus of the vagus nerve: a cyto- and chemoarchitectonic study in the human. J. Comp. Neurol., 330, 158–182.CrossRefGoogle ScholarPubMed
Huizinga, J. D., Thuneberg, L., Vanderwinden, J. M. and Rumessen, J. J. (1997). Interstitial cells of Cajal as targets for pharmacological intervention in gastrointestinal motor disorders. Trends Pharmacol. Sci., 18, 393–403.CrossRefGoogle ScholarPubMed
Huizinga, J. D., Ambrous, K. and Der-Silaphet, T. (1998). Co-operation between neural and myogenic mechanisms in the control of distension-induced peristalsis in the mouse small intestine. J. Physiol., 506, 843–856.CrossRefGoogle ScholarPubMed
Huizinga, J. D. and Faussone-Pellegrini, M. S. (2005). About the presence of interstitial cells of Cajal outside the musculature of the gastrointestinal tract. J. Cell Mol. Med., 9, 468–473.CrossRefGoogle ScholarPubMed
Hultborn, H. (2001). State-dependent modulation of sensory feedback. J. Physiol., 533, 5–13.CrossRefGoogle ScholarPubMed
Hummel, T., Sengupta, J. N., Meller, S. T. and Gebhart, G. F. (1997). Responses of T2–4 spinal cord neurons to irritation of the lower airways in the rat. Am. J. Physiol., 273, R1147–R1157.Google ScholarPubMed
Hunsperger, R. W. (1956). Affektreactionen auf elektrische Reizung im Hirnstamm der Katze [Affect reactions to electrical stimulation in the brain stem of the cat]. Helv. Physiol. Acta, 14, 70–92.Google Scholar
Iino, S., Ward, S. M. and Sanders, K. M. (2004). Interstitial cells of Cajal are functionally innervated by excitatory motor neurones in the murine intestine. J. Physiol., 556, 521–530.CrossRefGoogle ScholarPubMed
Inoue, T. (1980). Efferent discharge patterns in the ciliary nerve of rabbits and the pupillary light reflex. Brain Res., 186, 43–53.CrossRefGoogle ScholarPubMed
Ireland, D. R. (1999). Preferential formation of strong synapses during re-innervation of guinea-pig sympathetic ganglia. J. Physiol., 520, 827–837.CrossRefGoogle ScholarPubMed
Ireland, D. R., Davies, P. J. and McLachlan, E. M. (1999). Calcium channel subtypes differ at two types of cholinergic synapse in lumbar sympathetic neurones of guinea-pigs. J. Physiol., 514, 59–69.CrossRefGoogle ScholarPubMed
Ivanov, A. and Purves, D. (1989). Ongoing electrical activity of superior cervical ganglion cells in mammals of different size. J. Comp. Neurol., 284, 398–404.CrossRefGoogle ScholarPubMed
Iversen, S., Iversen, L. and Saper, C. B. (2000). The autonomic nervous system. In The Principles of Neural Science, 4th edn., ed. Kandel, E. R., Schwartz, J. H. and Jessel, T. M.. New York: McGraw-Hill, pp. 960–981.Google Scholar
Izumi, H. (1999). Nervous control of blood flow in the orofacial region. Pharmacol. Ther., 81, 141–161.CrossRefGoogle ScholarPubMed
Izzo, P. N. and Spyer, K. M. (1997). Parasympathetic innervation of the heart. In The Autonomic Nervous System, Vol. 11, Central Nervous Control of Autonomic function, ed. Jordan, D.. Amsterdam: Harwood Academic Publishers, pp. 109–127.Google Scholar
Izzo, P. N., Deuchars, J. and Spyer, K. M. (1993). Localization of cardiac vagal preganglionic motoneurones in the rat: immunocytochemical evidence of synaptic inputs containing 5- hydroxytryptamine. J. Comp. Neurol., 327, 572–583.CrossRefGoogle ScholarPubMed
Jack, J. B. B., Noble, D. and Tsien, R. W. (1975). Electrical Current Flow in Excitable Cells. Oxford: Clarendon Press.Google Scholar
James, W. (1884). What is an emotion?Mind, 9, 188–205.CrossRefGoogle Scholar
James, W. (1994) [1894]. The physical bases of emotion. 1894. Psychol. Rev., 101, 205–210.CrossRefGoogle ScholarPubMed
Jamieson, J., Boyd, H. D. and McLachlan, E. M. (2003). Simulations to derive membrane resistivity in three phenotypes of guinea pig sympathetic postganglionic neuron. J. Neurophysiol., 89, 2430–2440.CrossRefGoogle ScholarPubMed
Jan, L. Y. and Jan, Y. N. (1982). Peptidergic transmission in sympathetic ganglia of the frog. J. Physiol., 327, 219–246.CrossRefGoogle ScholarPubMed
Jan, Y. N., Jan, L. Y. and Kuffler, S. W. (1979). A peptide as a possible transmitter in sympathetic ganglia of the frog. Proc. Natl. Acad. Sci. U.S.A., 76, 1501–1505.CrossRefGoogle ScholarPubMed
Jan, L. Y., Jan, Y. N. and Brownfield, M. S. (1980a). Peptidergic transmitters in synaptic boutons of sympathetic ganglia. Nature, 288, 380–382.CrossRefGoogle Scholar
Jan, Y. N., Jan, L. Y. and Kuffler, S. W. (1980b). Further evidence for peptidergic transmission in sympathetic ganglia. Proc. Natl. Acad. Sci. U.S.A., 77, 5008–5012.CrossRefGoogle Scholar
Jancsó, N. (1960). Role of the nerve terminals in the mechanism of inflammatory reactions. Bull. Millard Fillmore Hosp. (Buffalo, NY), 7, 53–77.Google Scholar
Jancsó, N., Jancsó-Gábor, A. and Szolcsányi, J. (1967). Direct evidence for neurogenic inflammation and its prevention by denervation and by pretreatment with capsaicin. Br. J. Pharmacol., 31, 138–151.Google ScholarPubMed
Jancsó, N., Jancsó-Gábor, A. and Szolcsányi, J. (1968). The role of sensory nerve endings in neurogenic inflammation induced in human skin and in the eye and paw of the rat. Br. J. Pharmacol., 33, 32–41.Google ScholarPubMed
Jänig, W. (1975). Central organization of somatosympathetic reflexes in vasoconstrictor neurones. Brain Res., 87, 305–312.CrossRefGoogle ScholarPubMed
Jänig, W. (1985a). Organization of the lumbar sympathetic outflow to skeletal muscle and skin of the cat hindlimb and tail. Rev. Physiol. Biochem. Pharmacol., 102, 119–213.CrossRefGoogle Scholar
Jänig, W. (1985b). Causalgia and reflex sympathetic dystrophy: in which way is the sympathetic nervous system involved. Trends Neurosci., 8, 471–477.CrossRefGoogle Scholar
Jänig, W. (1986). Spinal cord integration of visceral sensory systems and sympathetic nervous system reflexes. Prog. Brain Res., 67, 255–277.CrossRefGoogle ScholarPubMed
Jänig, W. (1988a). Pre- and postganglionic vasoconstrictor neurons: differentiation, types, and discharge properties. Annu. Rev. Physiol., 50, 525–539.CrossRefGoogle Scholar
Jänig, W. (1988b). Integration of gut function by sympathetic reflexes. Bailliere's Clin. Gastroenterol., 2, 45–62.CrossRefGoogle Scholar
Jänig, W. (1990a). Functions of the sympathetic innervation of the skin. In Central Regulation of Autonomic Function, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 334–348.Google Scholar
Jänig, W. (1990b). The sympathetic nervous system in pain: physiology and pathophysiology. In Pain and the Sympathetic Nervous System, ed. Stanton-Hicks, M.. Boston, Dordrecht, London: Kluwer Academic Publishers, pp. 17–89.Google Scholar
Jänig, W. (1993). Spinal visceral afferents, sympathetic nervous system and referred pain. In New Trends in Referred Pain and Hyperalgesia, Pain Research and Clinical Management, Vol. 7, ed. Vecchiet, L., Albe-Fessard, D., Lindblom, U. and Giamberardino, M. A.. Amsterdam: Elsevier Science Publishers, pp. 83–92.Google Scholar
Jänig, W. (1995a). Ganglionic transmission in vivo. In The Autonomic Nervous System, Vol. 6, Autonomic Ganglia, ed. McLachlan, E. M.. Chur: Harwood Academic Publishers, pp. 349–395.Google Scholar
Jänig, W. (1995b). The sympathetic nervous system in pain. Eur. J. Anaesthesiol., 12 (Suppl. 10), 53–60.Google Scholar
Jänig, W. (1996a). Spinal cord reflex organization of sympathetic systems. Prog. Brain Res., 107, 43–77.CrossRefGoogle Scholar
Jänig, W. (1996b). Regulation of the lower urinary tract. In Comprehensive Human Physiology, Vol. 2, ed. Greger, R. and Windhorst, U.. Berlin, Heidelberg: Springer-Verlag, pp. 1611–1624.CrossRefGoogle Scholar
Jänig, W. (1996c). Behavioral and neurovegetative components of reproductive functions. In Comprehensive Human Physiology, Vol. 2, ed. Greger, R. and Windhorst, U.. Berlin, Heidelberg: Springer-Verlag, pp. 2253–2263.CrossRefGoogle Scholar
Jänig, W. (1996d). Neurobiology of visceral afferent neurons: neuroanatomy, functions, organ regulations and sensations. Biol. Psychol., 42, 29–51.CrossRefGoogle Scholar
Jänig, W. (2002). Pain in the sympathetic nervous system: pathophysiological mechanisms. In Autonomic Failure, 4th edn., ed. Bannister, R. and Mathias, C. J.. New York, Oxford: Oxford University Press, pp. 99–108.Google Scholar
Jänig, W. (2005a). Vegetatives Nervensystem. In Physiologie des Menschen, 29th edn., ed. Schmidt, R. F., Lang, F. and Thews, G.. Heidelberg Berlin: Springer Medizin Verlag, pp. 425–458.CrossRefGoogle Scholar
Jänig, W. (2005b). Vagal afferents and visceral pain. In Advances in Vagal Afferent Neurobiology, ed. Undem, B. and Weinreich, D.. Boca Raton: CRC Press, pp. 461–489.CrossRefGoogle Scholar
Jänig, W. and Baron, R. (2001). The role of the sympathetic nervous system in neuropathic pain: clinical observations and animal models. In Neuropathic Pain: Pathophysiological and Treatment, ed. Hansson, P. T., Fields, H. L., Hill, R. G. and Marchettini, P.. Seattle: IASP Press, pp. 125–149.Google Scholar
Jänig, W. and Baron, R. (2002). Complex regional pain syndrome is a disease of the central nervous system. Clin. Auton. Res., 12, 150–164.CrossRefGoogle ScholarPubMed
Jänig, W. and Baron, R. (2003). Complex regional pain syndrome: mystery explained? Lancet Neurol., 2, 687–697.CrossRefGoogle ScholarPubMed
Jänig, W. and Häbler, H. J. (1995). Visceral-autonomic integration. In Visceral Pain; Progress in Pain Research and Management, Vol. 5, ed. Gebhart, G. F.. Seattle: IASP Press, pp. 311–348.Google Scholar
Jänig, W. and Häbler, H. J. (1999). Organisation of the autonomic nervous system: structure and function. In Handbook of Clinical Neurology, Vol. 74, The Autonomic Nervous System, part I: Normal Functions, ed. Appenzeller, O.. Amsterdam: Elsevier, pp. 1–52.Google Scholar
Jänig, W. and Häbler, H. J. (2000a). Specificity in the organization of the autonomic nervous system: a basis for precise neural regulation of homeostatic and protective body functions. Prog. Brain Res., 122, 351–367.CrossRefGoogle Scholar
Jänig, W. and Häbler, H. J. (2000b). Sympathetic nervous system: contribution to chronic pain. Prog. Brain Res., 129, 451–468.Google Scholar
Jänig, W. and Häbler, H. J. (2002). Physiologie und pathophysiologie viszeraler Schmerzen [Physiology and pathophysiology of visceral pain]. Schmerz., 16, 429–446.CrossRefGoogle Scholar
Jänig, W. and Häbler, H. J. (2003). Neurophysiological analysis of target-related sympathetic pathways – from animal to human: similarities and differences. Acta Physiol. Scand., 177, 255–274.CrossRefGoogle ScholarPubMed
Jänig, W. and Koltzenburg, M. (1990). On the function of spinal primary afferent fibres supplying colon and urinary bladder. J. Auton. Nerv. Syst., 30 Suppl, S89–S96.CrossRefGoogle ScholarPubMed
Jänig, W. and Koltzenburg, M. (1991a). What is the interaction between the sympathetic terminal and the primary afferent fiber? In Towards a New Pharmacotherapy of Pain, ed. Basbaum, A. I. and Besson, J.-M.. Chichester: Dahlem Workshop Reports John Wiley & Sons, pp. 331–352.Google Scholar
Jänig, W. and Koltzenburg, M. (1991b). Plasticity of sympathetic reflex organization following cross-union of inappropriate nerves in the adult cat. J. Physiol., 436, 309–323.CrossRefGoogle Scholar
Jänig, W. and Koltzenburg, M. (1991c). Receptive properties of sacral primary afferent neurons supplying the colon. J. Neurophysiol., 65, 1067–1077.CrossRefGoogle Scholar
Jänig, W. and Koltzenburg, M. (1993). Pain arising from the urogenital tract. In The Autonomic Nervous System, Vol. 3., Nervous Control of the Urogenital Tract, ed. Maggi, C. A.. Chur, Switzerland: Harwood Academic Publisher, pp. 523–576.Google Scholar
Jänig, W. and Kümmel, H. (1977). Functional discrimination of postganglionic neurones to the cat's hindpaw with respect to the skin potentials recorded from the hairless skin. Pflügers Arch., 371, 217–225.CrossRefGoogle ScholarPubMed
Jänig, W. and Kümmel, H. (1981). Organization of the sympathetic innervation supplying the hairless skin of the cat's paw. J. Auton. Nerv. Syst., 3, 215–230.CrossRefGoogle ScholarPubMed
Jänig, W. and Levine, J. D. (2005). Autonomic-endocrine-immune responses in acute and chronic pain. In Wall and Melzack's Textbook of Pain, 5th edn., ed. McMahon, S. B. and Koltzenburg, M.. Edinburgh: Elsevier Churchill Livingstone, pp. 205–218.Google Scholar
Jänig, W. and Lisney, S. J. (1989). Small diameter myelinated afferents produce vasodilation but not plasma extravasation in rat skin. J. Physiol., 415, 477–486.CrossRefGoogle Scholar
Jänig, W. and McLachlan, E. M. (1986a). The sympathetic and sensory components of the caudal lumbar sympathetic trunk in the cat. J. Comp. Neurol., 245, 62–73.CrossRefGoogle Scholar
Jänig, W. and McLachlan, E. M. (1986b). Identification of distinct topographical distributions of lumbar sympathetic and sensory neurons projecting to end organs with different functions in the cat. J. Comp. Neurol., 246, 104–112.CrossRefGoogle Scholar
Jänig, W. and McLachlan, E. M. (1987). Organization of lumbar spinal outflow to distal colon and pelvic organs. Physiol. Rev., 67, 1332–1404.CrossRefGoogle ScholarPubMed
Jänig, W. and McLachlan, E. M. (1992a). Characteristics of function-specific pathways in the sympathetic nervous system. Trends Neurosci., 15, 475–481.CrossRefGoogle Scholar
Jänig, W. and McLachlan, E. M. (1992b). Specialized functional pathways are the building blocks of the autonomic nervous system. J. Auton. Nerv. Syst., 41, 3–13.CrossRefGoogle Scholar
Jänig, W. and McLachlan, E. M. (1994). The role of modifications in noradrenergic peripheral pathway after nerve lesions in the generation of pain. In Pharmacological Approaches to the Treatment of Pain: New Concepts and Critical Issues, Progress in Pain Research and Management, Vol. 1, ed. Fields, H. L. and Liebeskind, J. C.. Seattle: IASP Press, pp. 101–128.Google Scholar
Jänig, W. and McLachlan, E. M. (2002). Neurobiology of the autonomic nervous system. In Autonomic Failure, 4th edn., ed. Mathias, C. J. and Bannister, R.. New York, Oxford: Oxford University Press, pp. 3–15.Google ScholarPubMed
Jänig, W. and Morrison, J. F. B. (1986). Functional properties of spinal visceral afferents supplying abdominal and pelvic organs, with special emphasis on visceral nociception. Prog. Brain Res., 67, 87–114.CrossRefGoogle ScholarPubMed
Jänig, W. and Räth, B. (1977). Electrodermal reflexes in the cat's paws elicited by natural stimulation of skin. Pflügers Arch., 369, 27–32.CrossRefGoogle Scholar
Jänig, W. and Räth, B. (1980). Effects of anaesthetics on reflexes elicited in the sudomotor system by stimulation of Pacinian corpuscles and of cutaneous nociceptors. J. Auton. Nerv. Syst., 2, 1–14.CrossRefGoogle ScholarPubMed
Jänig, W. and Spilok, N. (1978). Functional organization of the sympathetic innervation supplying the hairless skin of the hindpaws in chronic spinal cats. Pflügers Arch., 377, 25–31.CrossRefGoogle ScholarPubMed
Jänig, W. and Stanton-Hicks, M. (eds.) (1996). Reflex Sympathetic Dystrophy – a Reappraisal. Seattle: IASP Press.Google Scholar
Jänig, W. and Szulczyk, P. (1980). Functional properties of lumbar preganglionic neurones. Brain Res., 186, 115–131.CrossRefGoogle ScholarPubMed
Jänig, W. and Szulczyk, P. (1981). The organization of lumbar preganglionic neurons. J. Auton. Nerv. Syst., 3, 177–191.CrossRefGoogle ScholarPubMed
Jänig, W., Krauspe, R. and Wiedersatz, G. (1982). Transmission of impulses from pre- to postganglionic vasoconstrictor and sudomotor neurons. J. Auton. Nerv. Syst., 6, 95–106.CrossRefGoogle ScholarPubMed
Jänig, W., Krauspe, R. and Wiedersatz, G. (1983). Reflex activation of postganglionic vasoconstrictor neurones supplying skeletal muscle by stimulation of arterial chemoreceptors via non-nicotinic synaptic mechanisms in sympathetic ganglia. Pflügers Arch., 396, 95–100.CrossRefGoogle ScholarPubMed
Jänig, W., Krauspe, R. and Wiedersatz, G. (1984). Activation of postganglionic neurones via non-nicotinic synaptic mechanisms by stimulation of thin preganglionic axons. Pflügers Arch., 401, 318–320.CrossRefGoogle ScholarPubMed
Jänig, W., Schmidt, M., Schnitzler, A. and Wesselmann, U. (1991). Differentiation of sympathetic neurones projecting in the hypogastric nerves in terms of their discharge patterns in cats. J. Physiol., 437, 157–179.CrossRefGoogle ScholarPubMed
Jänig, W., Levine, J. D. and Michaelis, M. (1996). Interactions of sympathetic and primary afferent neurons following nerve injury and tissue trauma. Prog. Brain Res., 112, 161–184.CrossRefGoogle Scholar
Jänig, W., Khasar, S. G., Levine, J. D. and Miao, F. J. P. (2000). The role of vagal visceral afferents in the control of nociception. Prog. Brain Res., 122, 273–287.CrossRefGoogle ScholarPubMed
Jänig, W., Chapman, C. R. and Green, P. G. (2006). Pain and body protection: sensory, autonomic, neuroendocrine and behavioral mechanisms in the control of inflammation and hyperalgesia. In Proceeding of the 11th World Congress on Pain, ed. Flor, H., Kalso, E. and Dostrovsky, J. O.. Seattle: IASP Press, in press.Google Scholar
Jankowska, E. (1992). Interneuronal relay in spinal pathways from proprioceptors. Prog. Neurobiol., 38, 335–378.CrossRefGoogle ScholarPubMed
Jankowska, E. (2001). Spinal interneuronal systems: identification, multifunctional character and reconfigurations in mammals. J. Physiol., 533, 31–40.CrossRefGoogle ScholarPubMed
Jankowska, E. and Lundberg, A. (1981). Interneurones in the spinal cord. Trends Neurosci., 4, 230–233.CrossRefGoogle Scholar
Jansen, A. S. and Loewy, A. D. (1997). Neurons lying in the white matter of the upper cervical spinal cord project to the intermediolateral cell column. Neuroscience, 77, 889–898.CrossRefGoogle ScholarPubMed
Jansen, A. S., Horst, G. J., Mettenleiter, T. C. and Loewy, A. D. (1992). central nervous system cell groups projecting to the submandibular parasympathetic preganglionic neurons in the rat: a retrograde transneuronal viral cell body labeling study. Brain Res., 572, 253–260.CrossRefGoogle ScholarPubMed
Jansen, A. S., Farwell, D. G. and Loewy, A. D. (1993). Specificity of pseudorabies virus as a retrograde marker of sympathetic preganglionic neurons: implications for transneuronal labeling studies. Brain Res., 617, 103–112.CrossRefGoogle ScholarPubMed
Jansen, A. S., Nguyen, X. V., Karpitskiy, V., Mettenleiter, T. C. and Loewy, A. D. (1995a). Central command neurons of the sympathetic nervous system: basis of the fight-or-flight response. Science, 270, 644–646.CrossRefGoogle Scholar
Jansen, A. S., Wessendorf, M. W. and Loewy, A. D. (1995b). Transneuronal labeling of central nervous system neuropeptide and monoamine neurons after pseudorabies virus injections into the stellate ganglion. Brain Res., 683, 1–24.CrossRefGoogle Scholar
Jeske, I., Morrison, S. F., Cravo, S. L. and Reis, D. J. (1993). Identification of baroreceptor reflex interneurons in the caudal ventrolateral medulla. Am. J. Physiol., 264, R169–R178.Google ScholarPubMed
Jewett, D. L. (1964). Activity of single efferent fibres in the cervical vagus nerve of the dog, with special reference to possible cardio-inhibitory fibres. J. Physiol., 175, 321–357.CrossRefGoogle ScholarPubMed
Jichia, D. and Frank, J. L. (2000). Spinal cord disease trauma and autonomic nervous system dysfunction. In Handbook of Clinical Neurology, Vol. 75, The Autonomic Nervous System part II: Dysfunctions, ed. Appenzeller, O., Amsterdam: Elsevier Science, pp. 567–587.Google Scholar
Jobling, P. and McLachlan, E. M. (1992). An electrophysiological study of responses evoked in isolated segments of rat tail artery during growth and maturation. J. Physiol., 454, 83–105.CrossRefGoogle ScholarPubMed
Jobling, P., McLachlan, E. M., Jänig, W. and Anderson, C. R. (1992). Electrophysiological responses in the rat tail artery during reinnervation following lesions of the sympathetic supply. J. Physiol., 454, 107–128.CrossRefGoogle ScholarPubMed
Jobling, P., Gibbins, I. L. and Morris, J. L. (2003). Functional organization of vasodilator neurons in pelvic ganglia of female guinea pigs: comparison with uterine motor neurons. J. Comp. Neurol., 459, 223–241.CrossRefGoogle ScholarPubMed
Johnson, C. D. and Gilbey, M. P. (1996). On the dominant rhythm in the discharges of single postganglionic sympathetic neurones innervating the rat tail artery. J. Physiol., 497, 241–259.CrossRefGoogle ScholarPubMed
Johnson, D. A. and Purves, D. (1981). Post-natal reduction of neural unit size in the rabbit ciliary ganglion. J. Physiol., 318, 143–159.CrossRefGoogle ScholarPubMed
Johnson, D. A. and Purves, D. (1983). Tonic and reflex synaptic activity recorded in ciliary ganglion cells of anaesthetized rabbits. J. Physiol., 339, 599–613.CrossRefGoogle ScholarPubMed
Jones, E. G. (2002). A pain in the thalamus. J. Pain, 3, 102–104.CrossRefGoogle ScholarPubMed
Jones, E. G. (2006). The Thalamus, 2nd edn., 2 volumes. Cambridge: Cambridge University Press.Google Scholar
Jones, J. F., Wang, Y. and Jordan, D. (1998). Activity of C fibre cardiac vagal efferents in anaesthetized cats and rats. J. Physiol., 507, 869–880.CrossRefGoogle ScholarPubMed
Jordan, D. (1997a). Central nervous control of the airways. In The Autonomic Nervous System, Vol. 11, Central Nervous Control of Autonomic Function, ed. Jordan, D.. Amsterdam: Harwood Academic Publishers, pp. 63–107.Google Scholar
Jordan, D. (ed.) (1997b). The Autonomic Nervous System, Vol. 11, Central Nervous Control of Autonomic Function, Amsterdam: Harwood Academic Press.
Joyner, M. J. and Dietz, N. M. (2003). Sympathetic vasodilation in human muscle. Acta Physiol. Scand., 177, 329–336.CrossRefGoogle ScholarPubMed
Joyner, M. J. and Halliwill, J. R. (2000). Sympathetic vasodilatation in human limbs. J. Physiol., 526, 471–480.Google ScholarPubMed
Julé, Y. and Szurszewski, J. H. (1983). Electrophysiology of neurones of the inferior mesenteric ganglion of the cat. J. Physiol., 344, 277–292.CrossRefGoogle ScholarPubMed
Juler, G. L. and Eltorai, I. M. (1985). The acute abdomen in spinal cord injury patients. Paraplegia, 23, 118–123.Google ScholarPubMed
Kalia, M. and Richter, D. (1985a). Morphology of physiologically identified slowly adapting lung stretch receptor afferents stained with intra-axonal horseradish peroxidase in the nucleus of the tractus solitarius of the cat. I. A light microscopic analysis. J. Comp. Neurol., 241, 503–520.CrossRefGoogle Scholar
Kalia, M. and Richter, D. (1985b). Morphology of physiologically identified slowly adapting lung stretch receptor afferents stained with intra-axonal horseradish peroxidase in the nucleus of the tractus solitarius of the cat. II. An ultrastructural analysis. J. Comp. Neurol., 241, 521–535.CrossRefGoogle Scholar
Kalia, M. and Richter, D. (1988a). Rapidly adapting pulmonary receptor afferents: I. Arborization in the nucleus of the tractus solitarius. J. Comp. Neurol., 274, 560–573.CrossRefGoogle Scholar
Kalia, M. and Richter, D. (1988b). Rapidly adapting pulmonary receptor afferents: II. Fine structure and synaptic organization of central terminal processes in the nucleus of the tractus solitarius. J. Comp. Neurol., 274, 574–594.CrossRefGoogle Scholar
Kanosue, K., Hosono, T., Zhang, Y. H. and Chen, X. M. (1998). Neuronal networks controlling thermoregulatory effectors. Prog. Brain Res., 115, 49–62.CrossRefGoogle Scholar
Karczmar, K., Koketsu, K. and Nishi, S. (eds.) (1986). Autonomic and Enteric Ganglia. New York: Plenum Press.CrossRefGoogle Scholar
Karlsson, A. K. (1999). Autonomic dysreflexia. Spinal Cord, 37, 383–391.CrossRefGoogle ScholarPubMed
Katayama, Y. and Nishi, S. (1986). Peptidergic transmission. In Autonomic and Enteric Ganglia, ed. Karczmar, K., Koketsu, K. and Nishi, S.. New York, London: Plenum Press, pp. 181–200.CrossRefGoogle Scholar
Katona, P. G., Poitras, J. W., Barnett, G. O. and Terry, B. S. (1970). Cardiac vagal efferent activity and heart period in the carotid sinus reflex. Am. J. Physiol., 218, 1030–1037.Google ScholarPubMed
Katter, J. T., Dado, R. J., Kostarczyk, E. and Giesler, G. J. Jr. (1996). Spinothalamic and spinohypothalamic tract neurons in the sacral spinal cord of rats. II. Responses to cutaneous and visceral stimuli. J. Neurophysiol., 75, 2606–2628.CrossRefGoogle ScholarPubMed
Kazuyuki, K., Hosono, T., Zhang, Y. H. and Chen, X. M. (1998). Neuronal networks controlling thermoregulatory effectors. Prog. Brain Res., 115, 49–62.Google ScholarPubMed
Keast, J. R. (1995a). Pelvic ganglia. In The Autonomic Nervous System, Vol. 6, Autonomic Ganglia, ed. McLachlan, E. M.. London: Harwood Academic Publishers GmbH, pp. 445–479.Google Scholar
Keast, J. R. (1995b). Visualization and immunohistochemical characterization of sympathetic and parasympathetic neurons in the male rat major pelvic ganglion. Neuroscience, 66, 655–662.CrossRefGoogle Scholar
Keast, J. R. (1999). Unusual autonomic ganglia: connections, chemistry, and plasticity of pelvic ganglia. Int. Rev. Cytol., 193, 1–69.CrossRefGoogle ScholarPubMed
Keast, J. R., McLachlan, E. M. and Meckler, R. L. (1993). Relation between electrophysiological class and neuropeptide content of guinea pig sympathetic prevertebral neurons. J. Neurophysiol., 69, 384–394.CrossRefGoogle ScholarPubMed
Keast, J. R., Luckensmeyer, G. B. and Schemann, M. (1995). All pelvic neurons in male rats contain immunoreactivity for the synthetic enzymes of either noradrenaline or acetylcholine. Neurosci. Lett., 196, 209–212.CrossRefGoogle ScholarPubMed
Keay, K. A. and Bandler, R. (2001). Parallel circuits mediating distinct emotional coping reactions to different types of stress. Neurosci. Biobehav. Rev., 25, 669–678.CrossRefGoogle ScholarPubMed
Keay, K. A. and Bandler, R. (2004). Periaqueductal gray. In The Rat Nervous System, 3rd edn., ed. Paxinos, G.. San Diego: Academic Press, pp. 243–257.Google Scholar
Keay, K. A., Clement, C. I., Owler, B., Depaulis, A. and Bandler, R. (1994). Convergence of deep somatic and visceral nociceptive information onto a discrete ventrolateral midbrain periaqueductal gray region. Neuroscience, 61, 727–732.CrossRefGoogle ScholarPubMed
Keay, K. A., Feil, K., Gordon, B. D., Herbert, H. and Bandler, R. (1997). Spinal afferents to functionally distinct periaqueductal gray columns in the rat: an anterograde and retrograde tracing study. J. Comp. Neurol., 385, 207–229.3.0.CO;2-5>CrossRefGoogle ScholarPubMed
Keay, K. A., Clement, C. I., Matar, W. M.et al. (2002). Noxious activation of spinal or vagal afferents evokes distinct patterns of fos-like immunoreactivity in the ventrolateral periaqueductal gray of unanaesthetised rats. Brain Res., 948, 122–130.CrossRefGoogle ScholarPubMed
Keef, K. D. and Kreulen, D. L. (1986). Venous mechanoreceptor input to neurones in the inferior mesenteric ganglion of the guinea-pig. J. Physiol., 377, 49–59.CrossRefGoogle ScholarPubMed
Kellogg, D. L. Jr., Johnson, J. M. and Kosiba, W. A. (1989). Selective abolition of adrenergic vasoconstrictor responses in skin by local iontophoresis of bretylium. Am. J. Physiol., 257, H1599–H1606.Google ScholarPubMed
Kellogg, D. L. Jr., Pergola, P. E., Piest, K. L.et al. (1995). Cutaneous active vasodilation in humans is mediated by cholinergic nerve cotransmission. Circ. Res., 77, 1222–1228.CrossRefGoogle ScholarPubMed
Kesler, B. S., Mazzone, S. B. and Canning, B. J. (2002). Nitric oxide-dependent modulation of smooth-muscle tone by airway parasympathetic nerves. Am. J. Respir. Crit. Care, 165, 481–488.CrossRefGoogle ScholarPubMed
Khasar, S. G., Green, P. G. and Levine, J. D. (1993). Comparison of intradermal and subcutaneous hyperalgesic effects of inflammatory mediators in the rat. Neurosci. Lett., 153, 215–218.CrossRefGoogle ScholarPubMed
Khasar, S. G., Miao, F. J. P. and Levine, J. D. (1995). Inflammation modulates the contribution of receptor-subtypes to bradykinin-induced hyperalgesia in the rat. Neuroscience, 69, 685–690.CrossRefGoogle ScholarPubMed
Khasar, S. G., Miao, F. J. P., Jänig, W. and Levine, J. D. (1998a). Modulation of bradykinin-induced mechanical hyperalgesia in the rat by activity in abdominal vagal afferents. Eur. J. Neurosci., 10, 435–444.CrossRefGoogle Scholar
Khasar, S. G., Miao, F. J. P., Jänig, W. and Levine, J. D. (1998b). Vagotomy-induced enhancement of mechanical hyperalgesia in the rat is sympathoadrenal-mediated. J. Neurosci., 18, 3043–3049.CrossRefGoogle Scholar
Khasar, S. G., Green, P. G., Miao, F. J. P. and Levine, J. D. (2003). Vagal modulation of nociception is mediated by adrenomedullary epinephrine in the rat. Eur. J. Neurosci., 17, 909–915.CrossRefGoogle ScholarPubMed
Kim, M., Chiego, D. J. Jr. and Bradley, R. M. (2004). Morphology of parasympathetic neurons innervating rat lingual salivary glands. Auton. Neurosci., 111, 27–36.CrossRefGoogle ScholarPubMed
Kirchheim, H. R., Just, A. and Ehmke, H. (1998). Physiology and pathophysiology of baroreceptor function and neuro-hormonal abnormalities in heart failure. Basic Res. Cardiol., 93, Suppl. 1, 1–22.CrossRefGoogle ScholarPubMed
Kirkup, A. J., Brunsden, A. M. and Grundy, D. (2001). Receptors and transmission in the brain-gut axis: potential for novel therapies. I. Receptors on visceral afferents. Am. J. Physiol. Gastrointest. Liver Physiol., 280, G787–G794.CrossRefGoogle ScholarPubMed
Klemm, M. F. (1995). Neuromuscular junctions made by nerve fibres supplying the longitudinal muscle of the guinea-pig ileum. J. Auton. Nerv. Syst., 55, 155–164.CrossRefGoogle ScholarPubMed
Klemm, M., Hirst, G. D. and Campbell, G. (1992). Structure of autonomic neuromuscular junctions in the sinus venosus of the toad. J. Auton. Nerv. Syst., 39, 139–150.CrossRefGoogle ScholarPubMed
Klemm, M. F., Helden, D. F. and Luff, S. E. (1993). Ultrastructural analysis of sympathetic neuromuscular junctions on mesenteric veins of the guinea pig. J. Comp. Neurol., 334, 159–167.CrossRefGoogle ScholarPubMed
Koepchen, H. P. (1962). Die Blutdruckrhythmik [The Rhythm of Blood Pressure]. Darmstadt: Dr. Dietrich Steinkopff.Google Scholar
Koepchen, H. P. (1983). Respiratory and cardiovascular “centres”: functional entirety or separate structures. In Central Environment and the Control Systems of Breathing and Circulation, ed. Schläfke, M. E., Koepchen, H. P. and See, W. R.. Berlin, Heidelberg, New York: Springer, pp. 221–237.CrossRefGoogle Scholar
Koepchen, H. P. and Thurau, K. (1959). Über die Entstehungsbedingungen der atemsynchronen Schwankungen des Vagustonus (respiratorische Arrhythmie) [On the origin of the respiratory synchronous changes of the vagus tone (respiratory arrhythmia)]. Pflügers Arch., 259, 10–30.CrossRefGoogle Scholar
Koepchen, H. P., Seller, H., Polster, J. and Langhorst, P. (1968). Über die Fein-Vasomotorik der Muskelstrombahn und ihre Beziehung zur Ateminnervation [Spontaneous vasomotor changes in the muscle and their relation to the respiratory rhythm]. Pflügers Arch., 302, 285–299.CrossRefGoogle Scholar
Koepchen, H. P., Klüssendorf, D. and Sommer, D. (1981). Neurophysiological background of central neural cardiovascular-respiratory coordination: basic remarks and experimental approach. J. Auton. Nerv. Syst., 3, 335–368.CrossRefGoogle ScholarPubMed
Koepchen, H. P., Abel, H.-H. and Klüssendorf, D. (1987). Brain stem generation of specific and non-specific rhythms. In Organization of the Automatic Nervous System: Central and Peripheral Mechanisms, ed. Ciriello, J., Calaresu, F. R., Renaud, L. P. and Polosa, C.. New York: Alan R. Liss., pp. 179–188.Google Scholar
Koh, S. D., Ward, S. M., Ordog, T., Sanders, K. M. and Horowitz, B. (2003). Conductances responsible for slow wave generation and propagation in interstitial cells of Cajal. Curr. Opin. Pharmacol., 3, 579–582.CrossRefGoogle ScholarPubMed
Koltzenburg, M., Häbler, H. J. and Jänig, W. (1995). Functional reinnervation of the vasculature of the adult cat paw pad by axons originally innervating vessels in hairy skin. Neuroscience, 67, 245–252.CrossRefGoogle ScholarPubMed
Kopin, I. J. (1989). Plasma levels of catecholamines and dopamine-beta-hydroxylase. In Handbook of Experimental Pharmacology, Vol. 90/II, Catecholamines II, ed. Trendelenburg, U. and Weiner, N.. Berlin: Springer-Verlag, pp. 211–275.Google Scholar
Kopp, U. C. and DiBona, G. F. (2000). The neural control of renal function. In The Kidney: Physiology and Pathophysiology, 3rd edn., ed. Seldin, G. and Giebisch, G.. New York: Raven Press, pp. 981–1006.Google Scholar
Korner, P. I. (1979). Central nervous control of autonomic cardiovascular function. In Handbook of Physiology, The Cardiovascular System, Vol. I, The Heart. ed. Barne, R. M.. Bethesda: American Physiological Society, pp. 691–739.Google Scholar
Koshiya, N. and Guyenet, P. G. (1996). Tonic sympathetic chemoreflex after blockade of respiratory rhythmogenesis in the rat. J. Physiol., 491, 859–869.CrossRefGoogle ScholarPubMed
Koshiya, N., Huangfu, D. and Guyenet, P. G. (1993). Ventrolateral medulla and sympathetic chemoreflex in the rat. Brain Res., 609, 174–184.CrossRefGoogle ScholarPubMed
Kostarczyk, E., Zhang, X. and Giesler, G. J. Jr. (1997). Spinohypothalamic tract neurons in the cervical enlargement of rats: locations of antidromically identified ascending axons and their collateral branches in the contralateral brain. J. Neurophysiol., 77, 435–451.CrossRefGoogle ScholarPubMed
Koushanpour, E. (1991). Baroreceptor discharge behavior and resetting. In Baroreceptor Reflexes, ed. Persson, P. B. and Kirchheim, H. R.. Berlin, Heidelberg: Springer-Verlag, pp. 9–44.CrossRefGoogle Scholar
Krassioukov, A. V., Johns, D. G. and Schramm, L. P. (2002). Sensitivity of sympathetically correlated spinal interneurons, renal sympathetic nerve activity, and arterial pressure to somatic and visceral stimuli after chronic spinal injury. J. Neurotrauma, 19, 1521–1529.CrossRefGoogle ScholarPubMed
Kreis, M. E., Jiang, W., Kirkup, A. J. and Grundy, D. (2002). Cosensitivity of vagal mucosal afferents to histamine and 5-hydroxytryptamine (serotonin) in the rat jejunum. Am. J. Physiol. Gastrointest. Liver Physiol., 283, G612–G617.CrossRefGoogle ScholarPubMed
Krenz, N. R. and Weaver, L. C. (2000). Nerve growth factor in glia and inflammatory cells of the injured rat spinal cord. J. Neurochem., 74, 730–739.CrossRefGoogle ScholarPubMed
Krenz, N. R., Meakin, S. O., Krassioukov, A. V. and Weaver, L. C. (1999). Neutralizing intraspinal nerve growth factor blocks autonomic dysreflexia caused by spinal cord injury. J. Neurosci., 19, 7405–7414.CrossRefGoogle ScholarPubMed
Kress, M., Koltzenburg, M., Reeh, P. W. and Handwerker, H. O. (1992). Responsiveness and functional attributes of electrically localized terminals of cutaneous C-fibers in vivo and in vitro. J. Neurophysiol., 68, 581–595.CrossRefGoogle ScholarPubMed
Kreulen, D. L. and Peters, S. (1986). Non-cholinergic transmission in a sympathetic ganglion of the guinea-pig elicited by colon distension. J. Physiol., 374, 315–334.CrossRefGoogle Scholar
Kreulen, D. L. and Szurszewski, J. H. (1979a). Nerve pathways in celiac plexus of the guinea pig. Am. J. Physiol., 237, E90–E97.Google Scholar
Kreulen, D. L. and Szurszewski, J. H. (1979b). Reflex pathways in the abdominal prevertebral ganglia: evidence for a colo-colonic inhibitory reflex. J. Physiol., 295, 21–32.CrossRefGoogle Scholar
Krier, J. and Hartman, D. A. (1984). Electrical properties and synaptic connections to neurons in parasympathetic colonic ganglia of the cat. Am. J. Physiol., 247, G52–G61.Google ScholarPubMed
Krier, J. and Szurszewski, J. H. (1982). Effect of substance P on colonic mechanoreceptors, motility, and sympathetic neurons. Am. J. Physiol., 243, G259–G267.Google ScholarPubMed
Krier, J., Schmalz, P. F. and Szurszewski, J. H. (1982). Central innervation of neurones in the inferior mesenteric ganglion and of the large intestine of the cat. J. Physiol., 332, 125–138.CrossRefGoogle ScholarPubMed
Kuhn, R. A. (1950). Functional capacity of the isolated human spinal cord. Brain, 73, 1–51.CrossRefGoogle ScholarPubMed
Kumazawa, T. (1986). Sensory innervation of reproductive organs. Prog. Brain Res., 67, 115–131.CrossRefGoogle ScholarPubMed
Kumazawa, T. (1990). Functions of the nociceptive primary neurons. Jpn. J. Physiol., 40, 1–14.CrossRefGoogle ScholarPubMed
Kumazawa, T., Mizumura, K. and Sato, J. (1987). Response properties of polymodal receptors studied using in vitro testis superior spermatic nerve preparations of dogs. J. Neurophysiol., 57, 702–711.CrossRefGoogle ScholarPubMed
Kümmel, H. (1983). Activity in sympathetic neurons supplying skin and skeletal muscle in spinal cats. J. Auton. Nerv. Syst., 7, 319–327.CrossRefGoogle ScholarPubMed
Kunitake, T. and Kannan, H. (2000). Discharge pattern of renal sympathetic nerve activity in the conscious rat: spectral analysis of integrated activity. J. Neurophysiol., 84, 2859–2867.CrossRefGoogle ScholarPubMed
Kuntz, A. (1940). The structural organization of the inferior mesenteric ganglia. J. Comp. Neurol., 72, 371–382.CrossRefGoogle Scholar
Kuntz, A. (1954). The Autonomic Nervous System. Philadelphia: Lea & Febinger.Google Scholar
Kuntz, A. and Saccomanno, G. (1944). Reflex inhibition of intestinal motility mediated through decentralized prevertebral ganglia. J. Neurophysiol., 7, 163–171.CrossRefGoogle Scholar
Kunze, D. L. (1972). Reflex discharge patterns of cardiac vagal efferent fibres. J. Physiol., 222, 1–15.CrossRefGoogle ScholarPubMed
Kunze, W. A. and Furness, J. B. (1999). The enteric nervous system and regulation of intestinal motility. Annu. Rev. Physiol., 61, 117–142.CrossRefGoogle ScholarPubMed
Kuo, D. C., Hisamitsu, T. and Groat, W. C. (1984). A sympathetic projection from sacral paravertebral ganglia to the pelvic nerve and to postganglionic nerves on the surface of the urinary bladder and large intestine of the cat. J. Comp. Neurol., 226, 76–86.CrossRefGoogle ScholarPubMed
LaBar, K. S. and LeDoux, J. E. (2001). Coping with danger: the neural basis of defensive behavior and fearful feelings. In Handbook of Physiology. Section 7: The Endocrine System, Vol. IV, Coping with the Environment: Neural and Neuroendocrine Mechanisms, ed. McEwen, B. S.. Oxford, New York: Oxford University Press, pp. 139–154.Google Scholar
Lal, S., Kirkup, A. J., Brunsden, A. M., Thompson, D. G. and Grundy, D. (2001). Vagal afferent responses to fatty acids of different chain length in the rat. Am. J. Physiol. Gastrointest. Liver Physiol., 281, G907–G915.CrossRefGoogle ScholarPubMed
Lamb, K., Kang, Y. M., Gebhart, G. F. and Bielefeldt, K. (2003). Gastric inflammation triggers hypersensitivity to acid in awake rats. Gastroenterology, 125, 1410–1418.CrossRefGoogle ScholarPubMed
Lammers, W. J. (2000). Propagation of individual spikes as “patches” of activation in isolated feline duodenum. Am. J. Physiol. Gastrointest. Liver Physiol., 278, G297–G307.CrossRefGoogle Scholar
LaMotte, R. H., Lundberg, L. E. R. and Torebjörk, H. E. (1992). Pain, hyperalgesia and activity in nociceptive C units in humans after intradermal injection of capsaicin. J. Physiol., 448, 749–764.CrossRefGoogle ScholarPubMed
Lange, C. S. (1920) [1887]. Über Gemüthsbewegungen [translated into English]. In The Emotions, ed. James, W. and Lange, C. G.. Baltimore: Williams and Wilkins.Google Scholar
Langley, J. N. (1891). On the course and connections of the secretory fibres supplying the sweat glands of the feet of the cat. J. Physiol., 12, 347–374.CrossRefGoogle ScholarPubMed
Langley, J. N. (1892). On the origin from the spinal cord of the cervical and upper thoracic sympathetic fibres, with some observations on white and grey rami communicantes. Phil. Trans. R. Soc. Lond. B, 183, 85–124.CrossRefGoogle Scholar
Langley, J. N. (1894a). The arrangement of the sympathetic nervous system, based chiefly on observation upon pilo-motor nerves. J. Physiol., 15, 176–244.CrossRefGoogle Scholar
Langley, J. N. (1894b). Further observations on the secretory and vaso-motor fibres of the foot of the cat, with notes on other sympathetic nerve fibres. J. Physiol., 17, 296–314.CrossRefGoogle Scholar
Langley, J. N. (1895). Note on the regeneration of prae-ganglionic fibres of the sympathetic. J. Physiol., 18, 80–84.CrossRefGoogle ScholarPubMed
Langley, J. N. (1897). On the regeneration of preganglionic and of postganglionic visceral nerve fibres. J. Physiol., 22, 215–230.CrossRefGoogle Scholar
Langley, J. N. (1900). The sympathetic and other related systems of nerves. In Textbook of Physiology, ed. Schäfer, E. A.. Edinburgh, London: Young J. Pentland, pp. 616–696.Google Scholar
Langley, J. N. (1903a). Das sympathische und verwandte nervöse System der Wirbeltiere (autonomes nervöses System) [The sympathetic and related nervous system of vertebrates (autonomic nervous system)]. Ergeb. Physiol., 27/II, 818–827.CrossRefGoogle Scholar
Langley, J. N. (1903b). The autonomic nervous system. Brain, 26, 1–26.CrossRefGoogle Scholar
Langley, J. N. (1921). The Autonomic Nervous System. Part I. Cambridge: W. Heffer.Google Scholar
Langley, J. N. and Anderson, H. K. (1895a). On the innervation of the pelvic and adjoining viscera. Part I. The lower portion of the intestine. J. Physiol., 18, 67–105.CrossRefGoogle Scholar
Langley, J. N. and Anderson, H. K. (1895b). The innervation of the pelvic and adjoining viscera. Part II. The bladder. J. Physiol., 19, 71–84.CrossRefGoogle Scholar
Langley, J. N. and Anderson, H. K. (1895c). The innervation of the pelvic and adjoining viscera. Part III. The external generative organs. J. Physiol., 19, 85–121.Google Scholar
Langley, J. N. and Anderson, H. K. (1895d). The innervation of the pelvic and adjoining viscera. Part IV. The internal generative organs. J. Physiol., 19, 122–130.Google Scholar
Langley, J. N. and Sherrington, C. S. (1891). On pilo-motor nerves. J. Physiol., 12, 278–291.CrossRefGoogle Scholar
Larsen, P. J., Enquist, L. W. and Card, J. P. (1998). Characterization of the multisynaptic neuronal control of the rat pineal gland using viral transneuronal tracing. Eur. J. Neurosci., 10, 128–145.CrossRefGoogle ScholarPubMed
Lavidis, N. A. and Bennett, M. R. (1992). Probabilistic secretion of quanta from visualized sympathetic nerve varicosities in mouse vas deferens. J. Physiol., 454, 9–26.CrossRefGoogle ScholarPubMed
Lawson, S. N. (1996). Neurochemistry of cutaneous nociceptors. In Neurobiology of Nociceptors, ed. Belmonte, C. and Cervero, F.. Oxford, New York, Tokyo: Oxford University Press, pp. 72–91.CrossRefGoogle Scholar
Lawson, S. N. (2005). The peripheral sensory nervous system: dorsal root ganglion neurons. In Peripheral Neuropathy, 4th edn., ed. Dyck, P. and Thomas, P. K.. Amsterdam: W. B. Saunders, Elsevier, pp. 163–202.Google Scholar
Douarin, N. M. and Kalcheim, C. (1999). The Neural Crest, 2nd edn. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
LeDoux, J. E. (1996). The Emotional Brain. New York: Simon & Shuster.Google Scholar
Lee, B. Y., Karmakar, M. G., Herz, B. L. and Sturgill, R. A. (1995). Autonomic dysreflexia revisited. J. Spinal Cord. Med., 18, 75–87.CrossRefGoogle ScholarPubMed
Lee, L. Y. and Pisarri, T. E. (2001). Afferent properties and reflex functions of bronchopulmonary C-fibers. Respir. Physiol., 125, 47–65.CrossRefGoogle ScholarPubMed
Leite-Panissi, C. R., Coimbra, N. C. and Menescal-de-Oliveira, L. (2003). The cholinergic stimulation of the central amygdala modifying the tonic immobility response and antinociception in guinea pigs depends on the ventrolateral periaqueductal gray. Brain Res. Bull., 60, 167–178.CrossRefGoogle ScholarPubMed
Levenson, R. W. (1993). Autonomic nervous system differences among emotions. Psychol. Sci., 3, 23–27.CrossRefGoogle Scholar
Levenson, R. W., Ekman, P. and Friesen, M. V. (1990). Voluntary facial action generates emotion-specific autonomic nervous system activity. Psychophysiology, 27, 363–384.CrossRefGoogle ScholarPubMed
Levenson, R. W., Carstensen, L. L., Friesen, W. V. and Ekman, P. (1991). Emotion, physiology, and expression in old age. Psychol. Aging, 6, 28–35.CrossRefGoogle ScholarPubMed
Levenson, R. W., Ekman, P., Heider, K. and Friesen, W. V. (1992). Emotion and autonomic nervous system activity in the Minangkabau of West Sumatra. J. Personal. Soc. Psychol., 62, 972–988.CrossRefGoogle ScholarPubMed
Levine, J. D., Fields, H. L. and Basbaum, A. L. (1993). Peptides and the primary afferent nociceptor. J. Neurosci., 13, 2273–2286.CrossRefGoogle ScholarPubMed
Lew, M. J., Rivers, R. J. and Duling, B. R. (1989). Arteriolar smooth muscle responses are modulated by an intramural diffusion barrier. Am. J. Physiol., 257, H10–H16.Google ScholarPubMed
Lewin, G. R. and McMahon, S. B. (1993). Muscle afferents innervating skin form somatotopically appropriate connections in the adult rat dorsal horn. Eur. J. Neurosci., 5, 1083–1092.CrossRefGoogle ScholarPubMed
Lewin, G. R., Ritter, A. M. and Mendell, L. M. (1993). Nerve growth factor-induced hyperalgesia in the neonatal and adult rat. J. Neurosci., 13, 2136–2148.CrossRefGoogle ScholarPubMed
Lewin, G. R., Rueff, A. and Mendell, L. M. (1994). Peripheral and central mechanisms of nerve growth factor-induced hyperalgesia. Eur. J. Neurosci., 6, 1903–1912.CrossRefGoogle Scholar
Lewis, D. I. and Coote, J. H. (1995). Chemical mediators of spinal inhibition of rat sympathetic neurones on stimulation in the nucleus tractus solitarii. J. Physiol., 486, 483–494.CrossRefGoogle ScholarPubMed
Lewis, D. I. and Coote, J. H. (1996). Baroreceptor-induced inhibition of sympathetic neurons by γ-aminobutyric acid acting at a spinal site. Am. J. Physiol., 270, H1885–H1892.Google Scholar
Lewis, M. W., Hermann, G. E., Rogers, R. C. and Travagli, R. A. (2002). In vitro and in vivo analysis of the effects of corticotropin releasing factor on rat dorsal vagal complex. J. Physiol., 543, 135–146.CrossRefGoogle ScholarPubMed
Li, Y. W. and Dampney, R. A. (1994). Expression of Fos-like protein in brain following sustained hypertension and hypotension in conscious rabbits. Neuroscience, 61, 613–634.CrossRefGoogle ScholarPubMed
Li, Y. W., Gieroba, Z. J., McAllen, R. M. and Blessing, W. W. (1991). Neurons in rabbit caudal ventrolateral medulla inhibit bulbospinal barosensitive neurons in rostral medulla. Am. J. Physiol., 261, R44–R51.Google ScholarPubMed
Lichtman, J. W. (1977). The reorganization of synaptic connexions in the rat submandibular ganglion during post-natal development. J. Physiol., 273, 155–177.CrossRefGoogle ScholarPubMed
Lichtman, J. W., Purves, D. and Yip, J. W. (1979). On the purpose of selective innervation of guinea-pig superior cervical ganglion cells. J. Physiol., 292, 69–84.CrossRefGoogle ScholarPubMed
Lichtman, J. W., Purves, D. and Yip, J. W. (1980). Innervation of sympathetic neurons in the guinea-pig thoracic chain. J. Physiol., 298, 285–299.CrossRefGoogle ScholarPubMed
Liddell, E. G. T. (1960). The Discovery of Reflexes. Oxford: Clarendon Press.Google Scholar
Lindgren, I. and Olivecrona, H. (1947). Surgical treatment of angina pectoris. J. Neurosurg., 4, 19–39.CrossRefGoogle ScholarPubMed
Lindh, B., Lundberg, J. M. and Hökfelt, T. (1989). neuropeptide Y-, galanin-, vasoactive intestinal peptide/PHI-, calcitonin gene-related peptide- and substance P-immunoreactive neuronal subpopulations in the cat autonomic and sensory ganglia and their projections. Cell Tissue Res., 256, 259–273.CrossRefGoogle ScholarPubMed
Lindh, B., Risling, M., Remahl, S., Terenius, L. and Hökfelt, T. (1993). Peptide-immunoreactive neurons and nerve in lumbosacral sympathetic ganglia: selective elimination of a pathway-specific expression of immunoreactivities following sciatic nerve resection in kittens. Neuroscience, 55, 545–562.CrossRefGoogle ScholarPubMed
Lipski, J., Kanjhan, R., Kruszewska, B. and Rong, W. (1996). Properties of presympathetic neurones in the rostral ventrolateral medulla in the rat: an intracellular study “in vivo'. J. Physiol., 490, 729–744.CrossRefGoogle ScholarPubMed
Lipski, J., Kawai, Y., Qi, J., Comer, A. and Win, J. (1998). Whole cell patch-clamp study of putative vasomotor neurons isolated from the rostral ventrolateral medulla. Am. J. Physiol., 274, R1099–R1110.Google ScholarPubMed
Lipski, J., Lin, J., Teo, M. Y. and Wyk, M. (2002). The network vs. pacemaker theory of the activity of RVL presympathetic neurons – a comparison with another putative pacemaker system. Auton. Neurosci., 98, 85–89.CrossRefGoogle Scholar
Llewellyn-Smith, I. J. and Weaver, L. C. (2001). Changes in synaptic inputs to sympathetic preganglionic neurons after spinal cord injury. J. Comp. Neurol., 435, 226–240.CrossRefGoogle ScholarPubMed
Llewellyn-Smith, I. J., Phend, K. D., Minson, J. B., Pilowsky, P. M. and Chalmers, J. P. (1992). Glutamate-immunoreactive synapses on retrogradely-labelled sympathetic preganglionic neurons in rat thoracic spinal cord. Brain Res., 581, 67–80.CrossRefGoogle ScholarPubMed
Llewellyn-Smith, I. J., Minson, J. B., Pilowsky, P. M., Arnolda, L. F. and Chalmers, J. P. (1995). The one hundred percent hypothesis: glutamate or γ-aminobutyric acid in synapses on sympathetic preganglionic neurons. Clin. Exp. Hypertens., 17, 323–333.CrossRefGoogle ScholarPubMed
Llewellyn-Smith, I. J., Cassam, A. K., Krenz, N. R., Krassioukov, A. V. and Weaver, L. C. (1997). Glutamate- and γ-aminobutyric acid-immunoreactive synapses on sympathetic preganglionic neurons caudal to a spinal cord transection in rats. Neuroscience, 80, 1225–1235.CrossRefGoogle Scholar
Llewellyn-Smith, I. J., Arnolda, L. F., Pilowsky, P. M., Chalmers, J. P. and Minson, J. B. (1998). γ-aminobutyric acid- and glutamate-immunoreactive synapses on sympathetic preganglionic neurons projecting to the superior cervical ganglion. J. Auton. Nerv. Syst., 71, 96–110.CrossRefGoogle ScholarPubMed
Lloyd, D. P. C. (1960). Spinal mechanisms involved in somatic activities. In Neurophysiology, Vol. II, ed. Field, J., Magoun, H. W. and Hall, V. E.. Washington: American Physiological Society, pp. 929–949.Google Scholar
Loewy, A. D. (1990a). Central autonomic pathways. In Central Regulation of Autonomic Functions, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 88–103.Google Scholar
Loewy, A. D. (1990b). Autonomic control of the eye. In Central Regulation of Autonomic Functions, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 268–285.Google Scholar
Loewy, A. D. (1998). Viruses as transneuronal tracers for defining neural circuits. Neurosci. Biobehav. Rev., 22, 679–684.CrossRefGoogle ScholarPubMed
Loewy, A. D. and Burton, H. (1978). Nuclei of the solitary tract: efferent projections to the lower brain stem and spinal cord of the cat. J. Comp. Neurol., 181, 421–449.CrossRefGoogle ScholarPubMed
Loewy, A. D. and Haxhiu, M. A. (1993). central nervous system cell groups projecting to pancreatic parasympathetic preganglionic neurons. Brain Res., 620, 323–330.CrossRefGoogle ScholarPubMed
Loewy, A. D. and Spyer, K. M. (1990a). Vagal preganglionic neurons. In Central Regulation of Autonomic Functions, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 68–87.Google Scholar
Loewy, A. D. and Spyer, K. M. (eds.) (1990b). Central Regulation of Autonomic Functions. New York, Oxford: Oxford University Press.Google Scholar
Loewy, A. D., Franklin, M. F. and Haxhiu, M. A. (1994). central nervous system monoamine cell groups projecting to pancreatic vagal motor neurons: a transneuronal labeling study using pseudorabies virus. Brain Res., 638, 248–260.CrossRefGoogle ScholarPubMed
Löfving, B. (1961). Cardiovascular adjustments induced from the rostral cingulate gyrus with special reference to sympatho-inhibitory mechanisms. Acta Physiol. Scand., 53 (Suppl184), 1–82.Google ScholarPubMed
Logan, S. D., Pickering, A. E., Gibson, I. C., Nolan, M. F. and Spanswick, D. (1996). Electrotonic coupling between rat sympathetic preganglionic neurones in vitro. J. Physiol., 495, 491–502.CrossRefGoogle ScholarPubMed
Lombardi, F., Della Bella, P., Casati, R. and Malliani, A. (1981). Effects of intracoronary administration of bradykinin on the impulse activity of afferent sympathetic unmyelinated fibers with left ventricular endings in the cat. Circ. Res., 48, 69–75.CrossRefGoogle ScholarPubMed
Longhurst, J. C. (1995). Chemosensitive abdominal visceral afferents. In Visceral Pain. Progress in Pain Research and Management, Vol. 5, ed. Gebhart, G. F.. Seattle: IASP Press, pp. 99–132.Google Scholar
Lovén, C. (1866). Über die Erweiterung von Arterien in Folge einer Nervenerregung [On the vasodilation of arteries as a consequence of nerve stimulation]. Ber. Verh. königl. -sächs. Ges. Wiss.: Math. -phys. Classe, 18, 85–110.Google Scholar
Lovick, T. A. (1987). Differential control of cardiac and vasomotor activity by neurons in nucleus paragigantocellularis lateralis in the cat. J. Physiol., 389, 23–35.CrossRefGoogle ScholarPubMed
Low, P. (ed.) (1993). Clinical Autonomic Disorders. Boston, Toronto, London: Brown and Company.Google Scholar
Luckensmeyer, G. B. and Keast, J. R. (1998a). Characterisation of the adventitial rectal ganglia in the male rat by their immunohistochemical features and projections. J. Comp. Neurol., 396, 429–441.3.0.CO;2-3>CrossRefGoogle Scholar
Luckensmeyer, G. B. and Keast, J. R. (1998b). Projections of pelvic autonomic neurons within the lower bowel of the male rat: an anterograde labelling study. Neuroscience, 84, 263–280.CrossRefGoogle Scholar
Luff, S. E. and McLachlan, E. M. (1989). Frequency of neuromuscular junctions on arteries of different dimensions in the rabbit, guinea pig and rat. Blood Vessels, 26, 95–106.Google ScholarPubMed
Luff, S. E., McLachlan, E. M. and Hirst, G. D. (1987). An ultrastructural analysis of the sympathetic neuromuscular junctions on arterioles of the submucosa of the guinea pig ileum. J. Comp. Neurol., 257, 578–594.CrossRefGoogle ScholarPubMed
Luff, S. E., Hengstberger, S. G., McLachlan, E. M. and Anderson, W. P. (1991). Two types of sympathetic axon innervating the juxtaglomerular arterioles of the rabbit and rat kidney differ structurally from those supplying other arteries. J. Neurocytol., 20, 781–795.CrossRefGoogle ScholarPubMed
Luff, S. E., Hengstberger, S. G., McLachlan, E. M. and Anderson, W. P. (1992). Distribution of sympathetic neuroeffector junctions in the juxtaglomerular region of the rabbit kidney. J Auton. Nerv. Syst., 40, 239–253.CrossRefGoogle ScholarPubMed
Luff, S. E., Young, S. B. and McLachlan, E. M. (1995). Proportions and structure of contacting and non-contacting varicosities in the perivascular plexus of the rat tail artery. J. Comp. Neurol., 361, 699–709.CrossRefGoogle ScholarPubMed
Luff, S. E., Young, S. B. and McLachlan, E. M. (2000). Ultrastructure of substance P-immunoreactive terminals and their relation to vascular smooth muscle cells of rat small mesenteric arteries. J. Comp. Neurol., 416, 277–290.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Lundberg, A. (1971). Function of the ventral spinocerebellar tract. A new hypothesis. Exp. Brain Res., 12, 317–330.Google ScholarPubMed
Lundberg, A. (1975). Control of spinal mechansims from the brain. In The Basic Neurosciences, ed. Brady, R. C.. New York: Raven, pp. 253–265.Google Scholar
Lundberg, A. (1979). Multisensory control of spinal reflex pathways. Prog. Brain Res., 50, 11–28.CrossRefGoogle ScholarPubMed
Lundberg, J. M. (1981). Evidence for coexistence of vasoactive intestinal polypeptide (vasoactive intestinal peptide) and acetylcholine in neurons of cat exocrine glands. Morphological, biochemical and functional studies. Acta Physiol. Scand., 496, 1–57.Google ScholarPubMed
Lundberg, J. M. (1996). Pharmacology of cotransmission in the autonomic nervous system: integrative aspects on amines, neuropeptides, adenosine triphosphate, amino acids and nitric oxide. Pharmacol. Rev., 48, 113–178.Google ScholarPubMed
Lundberg, J. M., Hemsen, A., Rudehill, A.et al. (1988). Neuropeptide Y- and alpha-adrenergic receptors in pig spleen: localization, binding characteristics, cyclic AMP effects and functional responses in control and denervated animals. Neuroscience, 24, 659–672.CrossRefGoogle ScholarPubMed
Lundgren, O. (1988). Nervous control of intestinal transport. In Bailliere's Clinical Gastroenterology, Vol. 2/1, Gastrointestinal Neurophysiology, ed. Grundy, D. and Read, N. W.. London: Balliere Tindall, pp. 85–106.Google Scholar
Lundgren, O. (1989). Enteric nervous control of mucosal functions of the small intestine in vivo. In Nerves and the Gastrointestinal Tract, ed. Singer, M. V. and Goebell, A.. Dordrecht, The Netherlands: Kluwer Academic Publishers, pp. 275–285.Google Scholar
Lundgren, O. (2000). Sympathetic input into the enteric nervous system. Gut, 47, Suppl. 4, iv33–iv35.CrossRefGoogle ScholarPubMed
Luo, M., Hess, M. C., Fink, G. D.et al. (2003). Differential alterations in sympathetic neurotransmission in mesenteric arteries and veins in DOCA-salt hypertensive rats. Auton. Neurosci., 104, 47–57.CrossRefGoogle ScholarPubMed
Lykken, D. T. (1998). A Tremor in the Blood. New York: Plenum Press.Google Scholar
Lynn, B. (1996a). Neurogenic inflammation caused by cutaneous polymodal receptors. Prog. Brain Res., 113, 361–368.CrossRefGoogle Scholar
Lynn, B. (1996b). Efferent function of nociceptors. In Neurobiology of Nociceptors, ed. Belmonte, C. and Cervero, F.. Oxford, New York, Tokyo: Oxford University Press, pp. 418–438.CrossRefGoogle Scholar
Lynn, B., Schütterle, S. and Pierau, F. K. (1996). The vasodilator component of neurogenic inflammation is caused by a special subclass of heat-sensitive nociceptors in the skin of the pig. J. Physiol., 494, 587–593.CrossRefGoogle ScholarPubMed
Lynn, P. A. and Blackshaw, L. A. (1999). In vitro recordings of afferent fibres with receptive fields in the serosa, muscle and mucosa of rat colon. J. Physiol., 518, 271–282.CrossRefGoogle ScholarPubMed
Lynn, P. A., Olsson, C., Zagorodnyuk, V., Costa, M. and Brookes, S. J. (2003). Rectal intraganglionic laminar endings are transduction sites of extrinsic mechanoreceptors in the guinea pig rectum. Gastroenterology, 125, 786–794.CrossRefGoogle ScholarPubMed
Macefield, V. G. and Wallin, B. G. (1995). Modulation of muscle sympathetic activity during spontaneous and artificial ventilation and apnoea in humans. J. Auton. Nerv. Syst., 53, 137–147.CrossRefGoogle ScholarPubMed
Macefield, V. G. and Wallin, B. G. (1996). The discharge behaviour of single sympathetic neurones supplying human sweat glands. J. Physiol., 61, 277–286.Google ScholarPubMed
Macefield, V. G. and Wallin, B. G. (1999a). Firing properties of single vasoconstrictor neurones in human subjects with high level of muscle sympathetic activity. J. Physiol., 516, 293–301.CrossRefGoogle Scholar
Macefield, V. G. and Wallin, B. G. (1999b). Respiratory and cardiac modulation of single sympathetic vasoconstrictor and sudomotor neurones to human skin. J. Physiol., 516, 303–314.CrossRefGoogle Scholar
Macefield, V. G., Wallin, B. G. and Vallbo, A. B. (1994). The discharge behaviour of single vasoconstrictor motoneurones in human muscle nerves. J. Physiol., 481, 799–809.CrossRefGoogle ScholarPubMed
Macefield, V. G., Elam, M. and Wallin, B. G. (2002). Firing properties of single postganglionic sympathetic neurons recorded in awake human subjects. Auton. Neurosci., 95, 146–159.CrossRefGoogle ScholarPubMed
Macefield, V. G., Sverrisdottir, Y. B. and Wallin, B. G. (2003). Resting discharge of human muscle spindles is not modulated by increases in sympathetic drive. J. Physiol., 551, 1005–1011.CrossRefGoogle Scholar
Madden, K. S. and Felten, D. L. (1995). Experimental basis for neural-immune interactions. Physiol. Rev., 75, 77–106.CrossRefGoogle ScholarPubMed
Madden, C. J. and Sved, A. F. (2003). Cardiovascular regulation after destruction of the C1 cell group of the rostral ventrolateral medulla in rats. Am. J. Physiol. Heart Circ. Physiol., 285, H2734–H2748.CrossRefGoogle ScholarPubMed
Madden, K. S., Sanders, V. M. and Felten, D. L. (1995). Catecholamine influences and sympathetic modulation of immune responsiveness. Rev. Pharmacol. Toxicol., 35, 417–448.CrossRefGoogle ScholarPubMed
Maggi, C. A. (ed.) (1993). The Autonomic Nervous System, Vol. 3, Nervous Control of the Urogenital System. Chur, Switzerland: Harwood Academic Publishers.Google Scholar
Maggi, C. A. and Meli, A. (1988). The sensory-efferent function of capsaicin-sensitive sensory neurons. Gen. Pharmacol., 19, 1–43.CrossRefGoogle ScholarPubMed
Maggi, C. A., Giachetti, A., Dey, R. D. and Said, S. I. (1995). Neuropeptides as regulators for airway function: vasoactive intestinal peptide and the tachykinins. Physiol. Rev., 75, 277–322.CrossRefGoogle ScholarPubMed
Maier, S. F. and Watkins, L. R. (1998). Cytokines for psychologists: implications of bidirectional immune-to-brain communication for understanding behavior, mood, and cognition. Psychol. Rev., 105, 83–107.CrossRefGoogle Scholar
Malliani, A. (1982). Cardiovascular sympathetic afferent fibers. Rev. Physiol. Biochem. Pharmacol., 94, 11–74.CrossRefGoogle Scholar
Malliani, A. and Lombardi, F. (1982). Consideration of the fundamental mechanisms eliciting cardiac pain. Am. Heart J., 103, 575–578.CrossRefGoogle ScholarPubMed
Malpas, S. C. (1998). The rhythmicity of sympathetic nerve activity. Prog. Neurobiol., 56, 65–96.CrossRefGoogle ScholarPubMed
Mancia, G., Baccelli, G. and Zanchetti, A. (1972). Hemodynamic responses to different emotional stimuli in the cat: patterns and mechanisms. Am. J. Physiol., 223, 925–933.Google ScholarPubMed
Mano, T. (1999). Muscular and cutaneous sympathetic activity. In Handbook of Clinical Neurology, Vol. 74, The Autonomic Nervous System, part I: Normal Functions, ed. Appenzeller, O.. Amsterdam: Elsevier, pp. 649–665.Google Scholar
Mantyh, P. W., Rogers, S. D., Honore, P.et al. (1997). Inhibition of hyperalgesia by ablation of lamina I spinal neurons expressing the substance P receptor. Science, 278, 275–279.CrossRefGoogle ScholarPubMed
Markakis, E. A. and Swanson, L. W. (1997). Spatiotemporal patterns of secretomotor neuron generation in the parvicellular neuroendocrine system. Brain Res. Brain Res. Rev., 24, 255–291.CrossRefGoogle ScholarPubMed
Marsh, D. R. and Weaver, L. C. (2004). Autonomic dysreflexia, induced by noxious or innocuous stimulation, does not depend on changes in dorsal horn substance p. J. Neurotrauma, 21, 817–828.CrossRefGoogle Scholar
Marsh, D. R., Wong, S. T., Meakin, S. O.et al. (2002). Neutralizing intraspinal nerve growth factor with a trkA-IgG fusion protein blocks the development of autonomic dysreflexia in a clip-compression model of spinal cord injury. J. Neurotrauma, 19, 1531–1541.CrossRefGoogle Scholar
Marshall, J. M. (1994). Peripheral chemoreceptors and cardiovascular regulation. Physiol. Rev., 74, 543–594.CrossRefGoogle ScholarPubMed
Marson, L. (1995). Central nervous system neurons identified after injection of pseudorabies virus into the rat clitoris. Neurosci. Lett., 190, 41–44.CrossRefGoogle ScholarPubMed
Marson, L. (1997). Identification of central nervous system neurons that innervate the bladder body, bladder base, or external urethral sphincter of female rats: a transneuronal tracing study using pseudorabies virus. J. Comp. Neurol., 389, 584–602.3.0.CO;2-X>CrossRefGoogle ScholarPubMed
Marson, L. and McKenna, K. E. (1996). central nervous system cell groups involved in the control of the ischiocavernosus and bulbospongiosus muscles: a transneuronal tracing study using pseudorabies virus. J. Comp. Neurol., 374, 161–179.3.0.CO;2-0>CrossRefGoogle ScholarPubMed
Marson, L., Platt, K. B. and McKenna, K. E. (1993). Central nervous system innervation of the penis as revealed by the transneuronal transport of pseudorabies virus. Neuroscience, 55, 263–280.CrossRefGoogle ScholarPubMed
Martinez-Pena y Valencuela, I., Rogers, R. C., Hermann, G. E. and Travagli, R. A. (2004). Norepinephrine effects on identified neurons of the rat dorsal motor nucleus of the vagus. Am. J. Physiol. Gastrointest. Liver Physiol., 286, G333–G339.Google Scholar
Martínez-Pinna, J., Davies, P. J. and McLachlan, E. M. (2000). Diversity of channels involved in Ca2 + activation of K+ channels during the prolonged AHP in guinea-pig sympathetic neurons. J. Neurophysiol., 84, 1346–1354.CrossRefGoogle Scholar
Mason, P. (2001). Contributions of the medullary raphe and ventromedial reticular region to pain modulation and other homeostatic functions. Annu. Rev. Neurosci., 24, 737–777.CrossRefGoogle ScholarPubMed
Mathias, C. J. and Bannister, R. (eds.) (2002). Autonomic Failure, 4th edn. Oxford: Oxford University Press.Google ScholarPubMed
Mathias, C. J. and Frankel, H. L. (2002). Autonomic disturbances and spinal cord lesions. In Autonomic Failure, 4th edn., ed. Mathias, C. J. and Bannister, R.. Oxford: Oxford University Press, pp. 494–513.Google ScholarPubMed
Matsuo, R. and Kang, Y. (1998). Two types of parasympathetic preganglionic neurones in the superior salivatory nucleus characterized electrophysiologically in slice preparations of neonatal rats. J. Physiol., 513, 157–170.CrossRefGoogle ScholarPubMed
Matsuo, R. and Yamamoto, T. (1989). Gustatory-salivary reflex: neural activity of sympathetic and parasympathetic fibers innervating the submandibular gland of the hamster. J. Auton. Nerv. Syst., 26, 187–197.CrossRefGoogle ScholarPubMed
Matsuo, R., Morimoto, T. and Kang, Y. (1998). Neural activity of the superior salivatory nucleus in rats. Eur. J. Morphol., 36 Suppl, 203–207.Google ScholarPubMed
Matthews, L. H. (1969). The Life of Mammals. London: Weidenfeld and Nicolson.Google Scholar
Matthews, M. R. and Cuello, A. C. (1984). The origin and possible significance of substance P immunoreactive networks in the prevertebral ganglia and related structures in the guinea-pig. Philos. Trans. R. Soc. London B Biol. Sci., 306, 247–276.CrossRefGoogle ScholarPubMed
Matthews, M. R., Connaughton, M. and Cuello, A. C. (1987). Ultrastructure and distribution of substance P-immunoreactive sensory collaterals in the guinea pig prevertebral sympathetic ganglia. J. Comp. Neurol., 258, 28–51.CrossRefGoogle ScholarPubMed
Mawe, G. M. (1995). Prevertebral, pancreatic and gallbladder ganglia: non-enteric ganglia that are involved in gastrointestinal function. In The Autonomic Nervous System, Vol. 6, Autonomic Ganglia, ed. McLachlan, E. M.. Luxembourg: Harwood Academic Publishers, pp. 397–444.Google Scholar
Mawe, G. M. (1998). Nerves and hormones interact to control gallbladder function. News Physiol. Sci., 13, 84–90.Google ScholarPubMed
Mayer, E. A. and Raybould, H. E. (eds.) (1993). Basic and Clinical Aspects of Chronic Abdominal Pain. Amsterdam: Elsevier Science Publishers B.V.Google Scholar
Mayer, E. A., Munakata, J., Mertz, H., Lembo, T. and Bernstein, C. N. (1995). Visceral hyperalgesia and irritable bowel syndrome. In Visceral Pain, ed. Gebhart, G. F.. Seattle: IASP Press, pp. 429–468.Google Scholar
McAllen, R. M. (1987). Central respiratory modulation of subretrofacial bulbospinal neurones in the cat. J. Physiol., 388, 533–545.CrossRefGoogle ScholarPubMed
McAllen, R. M. (1992). Actions of carotid chemoreceptors on subretrofacial bulbospinal neurons in the cat. J. Auton. Nerv. Syst., 40, 181–188.CrossRefGoogle ScholarPubMed
McAllen, R. M. and May, C. N. (1994a). Differential drives from rostral ventrolateral medullary neurons to three identified sympathetic outflows. Am. J. Physiol., 267, R935–R944.Google Scholar
McAllen, R. M. and May, C. N. (1994b). Effects of preoptic warming on subretrofacial and cutaneous vasoconstrictor neurons in anaesthetized cats. J. Physiol., 481, 719–730.CrossRefGoogle Scholar
McAllen, R. M. and Spyer, K. M. (1978a). Two types of vagal preganglionic motoneurones projecting to the heart and lungs. J. Physiol., 282, 353–364.CrossRefGoogle Scholar
McAllen, R. M. and Spyer, K. M. (1978b). The baroreceptor input to cardiac vagal motoneurones. J. Physiol., 282, 365–374.CrossRefGoogle Scholar
McAllen, R. M., Häbler, H. J., Michaelis, M., Peters, O. and Jänig, W. (1994). Monosynaptic excitation of preganglionic vasomotor neurons by subretrofacial neurons of the rostral ventrolateral medulla. Brain Res., 634, 227–234.CrossRefGoogle ScholarPubMed
McAllen, R. M., May, C. N. and Shafton, A. D. (1995). Functional anatomy of sympathetic premotor cell groups in the medulla. Clin. Exp. Hypertens., 17, 209–221.CrossRefGoogle ScholarPubMed
McAllen, R. M., May, C. N. and Campos, R. R. (1997). The supply of vasomotor drive to individual classes of sympathetic neuron. Clin. Exp. Hypertens., 19, 607–618.CrossRefGoogle ScholarPubMed
McCabe, P. M., Duan, Y. F., Winters, R. W.et al. (1994). Comparison of peripheral blood flow patterns associated with the defense reaction and the vigilance reaction in rabbits. Physiol. Behav., 56, 1101–1106.CrossRefGoogle ScholarPubMed
McCall, R. B., Gebber, G. L. and Barman, S. M. (1977). Spinal interneurons in the baroreceptor reflex arc. Am. J. Physiol., 232, H657–H665.Google ScholarPubMed
McCrea, D. A. (1994). Can sense be made of spinal interneuron circuits? In Movement Control, ed. Cordo, P. and Harnad, S.. Cambridge: Cambridge University Press, pp. 31–41.CrossRefGoogle Scholar
McCrea, D. A. (2001). Spinal circuitry of sensorimotor control of locomotion. J. Physiol., 533, 41–50.CrossRefGoogle ScholarPubMed
McDonald, D. M. (1990). The ultrastructure and permeability of tracheobronchial blood vessels in health and disease. Eur. Respir. J. Suppl., 12, 572s–585s.Google ScholarPubMed
McDonald, D. M. (1997). Neurogenic inflammation in the airways. In The Autonomic Nervous System, Vol. 7, Autonomic Control of the Respiratory System, ed. Barnes, P. J.. Amsterdam: Harwood Academic Publishers GmbH, pp. 249–289.Google Scholar
McDonald, D. M., Mitchell, R. A., Gabella, G. and Haskell, A. (1988). Neurogenic inflammation in the rat trachea. II. Identity and distribution of nerves mediating the increase in vascular permeability. J. Neurocytol., 17, 605–628.CrossRefGoogle ScholarPubMed
McEwen, B. S. (1998). Protective and damaging effects of stress mediators. New Engl. J. Med., 338, 171–179.CrossRefGoogle ScholarPubMed
McEwen, B. S. (2000). Protective and damaging effects of stress mediators: central role of the brain. Prog. Brain Res., 122, 25–34.CrossRefGoogle Scholar
McEwen, B. S. (ed.) (2001a). Handbook of Physiology. Section 7: The Endocrine System, Vol. IV, Coping with the Environment: Neural and Neuroendocrine Mechanisms. Oxford, New York: Oxford University Press.Google Scholar
McEwen, B. S. (2001b). Neurobiology of interpreting and responding to stressful events: paradigmatic role of the hippocampus. In Handbook of Physiology. Section 7: The Endocrine System, Vol. IV, Coping with the Environment: Neural and Neuroendocrine Mechanisms, ed. McEwen, B. S., pp. 155–178. Oxford University Press, Oxford New York.Google Scholar
McEwen, B. S. and Wingfield, J. C. (2003). The concept of allostasis in biology and biomedicine. Horm. Behav., 43, 2–15.CrossRefGoogle ScholarPubMed
McGaugh, J. L. (2000). Memory – a century of consolidation. Science, 287, 248–251.CrossRefGoogle ScholarPubMed
McGaugh, J. L. and Roozendaal, B. (2002). Role of adrenal stress hormones in forming lasting memories in the brain. Curr. Opin. Neurobiol., 12, 205–210.CrossRefGoogle Scholar
McKenna, K. E. (1999). Central nervous system pathways involved in the control of penile erection. Annu. Rev. Sex Res., 10, 157–183.Google ScholarPubMed
McKenna, K. E. (2000). The neural control of female sexual function. NeuroRehabilitation, 15, 133–143.Google ScholarPubMed
McKenna, K. E. (2001). Neural circuitry involved in sexual function. J. Spinal Cord Med., 24, 148–154.CrossRefGoogle ScholarPubMed
McKenna, K. E. (2002). The neurophysiology of female sexual function. World J. Urol., 20, 93–100.CrossRefGoogle ScholarPubMed
McKenna, K. E. and Marson, L. (1997). Spinal and brain stem control of sexual function. In The Autonomic Nervous System, Vol. 11, Central Nervous Control of Autonomic Function, ed. Jordan, D.. Amsterdam: Harwood Academic Publishers, pp. 151–187.Google Scholar
McKenna, K. E., Chung, S. K. and McVary, K. T. (1991). A model for the study of sexual function in anesthetized male and female rats. Am. J. Physiol., 261, R1276–R1285.Google Scholar
McKinley, M. J., Clarke, I. J. and Oldfield, B. J. (2004). Circumventricular organs. In The Human Nervous System, ed. Paxinos, G. and Mai, J. K.. Amsterdam: Elsevier Academic Press, pp. 562–591.Google Scholar
McLachlan, E. M. (1975). An analysis of the release of acetylcholine from preganglionic nerve terminals. J. Physiol., 245, 447–466.CrossRefGoogle ScholarPubMed
McLachlan, E. M. (1985). The components of the hypogastric nerve in male and female guinea pigs. J. Auton. Nerv. Syst., 13, 327–342.CrossRefGoogle ScholarPubMed
McLachlan, E. M. (ed.) (1995). The Autonomic Nervous System, Vol. 6, Autonomic Ganglia. Luxembourg: Harwood Academic Publishers.Google Scholar
McLachlan, E. M. and Hirst, G. D. (1980). Some properties of preganglionic neurons in upper thoracic spinal cord of the cat. J. Neurophysiol., 43, 1251–1265.CrossRefGoogle ScholarPubMed
McLachlan, E. M. and Jänig, W. (1983). The cell bodies of origin of sympathetic and sensory axons in some skin and muscle nerves of the cat hindlimb. J. Comp. Neurol., 214, 115–130.CrossRefGoogle ScholarPubMed
McLachlan, E. M. and Meckler, R. L. (1989). Characteristics of synaptic input to three classes of sympathetic neurone in the coeliac ganglion of the guinea-pig. J. Physiol., 415, 109–129.CrossRefGoogle ScholarPubMed
McLachlan, E. M., Davies, P. J., Häbler, H. J. and Jamieson, J. (1997). Ongoing and reflex synaptic events in rat superior cervical ganglion cells. J. Physiol., 501, 165–181.CrossRefGoogle Scholar
McLachlan, E. M., Häbler, H. J., Jamieson, J. and Davies, P. J. (1998). Analysis of the periodicity of synaptic events in neurones in the superior cervical ganglion of anaesthetized rats. J. Physiol., 511, 461–478.CrossRefGoogle ScholarPubMed
McMahon, S. B. (1996). nerve growth factor as a mediator of inflammatory pain. Philos. Trans. R. Soc. London B Biol. Sci., 351, 431–440.CrossRefGoogle ScholarPubMed
McMahon, S. B. and Morrison, J. F. (1982a). Spinal neurones with long projections activated from the abdominal viscera of the cat. J. Physiol., 322, 1–20.CrossRefGoogle Scholar
McMahon, S. B. and Morrison, J. F. (1982b). Two groups of spinal interneurones that respond to stimulation of the abdominal viscera of the cat. J. Physiol., 322, 21–34.CrossRefGoogle Scholar
Meckler, R. L. and Weaver, L. C. (1988). Characteristics of ongoing and reflex discharge of single splenic and renal sympathetic postganglionic fibres in the cat. J. Physiol., 396, 139–153.CrossRefGoogle Scholar
Mei, N. (1983). Sensory structures in the viscera. Prog. Sensory Physiol., 4, 1–42.CrossRefGoogle Scholar
Mei, N. (1985). Intestinal chemosensitivity. Physiol. Rev., 65, 211–237.CrossRefGoogle ScholarPubMed
Meller, S. T. and Gebhart, G. F. (1992). A critical review of the afferent pathways and the potential chemical mediators involved in cardiac pain. Neuroscience, 48, 501–524.CrossRefGoogle ScholarPubMed
Melnitchenko, L. V. and Skok, V. I. (1970). Natural electrical activity in mammalian parasympathetic ganglion neurones. Brain Res., 23, 277–279.CrossRefGoogle ScholarPubMed
Menendez, L., Bester, H., Besson, J. M. and Bernard, J. F. (1996). Parabrachial area: electrophysiological evidence for an involvement in cold nociception. J. Neurophysiol., 75, 2099–2116.CrossRefGoogle ScholarPubMed
Mense, S. and Craig, A. D. Jr. (1988). Spinal and supraspinal terminations of primary afferent fibers from the gastrocnemius-soleus muscle in the cat. Neuroscience, 26, 1023–1035.CrossRefGoogle ScholarPubMed
Messenger, J. P., Anderson, R. L. and Gibbins, I. L. (1999). Neurokinin-1 receptor localisation in guinea pig autonomic ganglia. J. Comp. Neurol., 412, 693–704.3.0.CO;2-T>CrossRefGoogle ScholarPubMed
Meyer, R. A., Davis, K. D., Cohen, R. H., Treede, R. D. and Campbell, J. N. (1991). Mechanically insensitive afferents (MIAs) in cutaneous nerves of monkey. Brain Res., 561, 252–261.CrossRefGoogle ScholarPubMed
Meyer, R. A., Ringkamp, M., Campbell, J. N. and Raja, S. N. (2005). Peripheral mechanisms of cutaneous nociception. In Wall and Mezack's Textbook of Pain, 5th edn., ed. McMahon, S. B. and Koltzenburg, M.. Edinburgh: Elsevier Churchill Livingstone, pp. 3–34.Google Scholar
Meyers, G. E. (1986). William James, His Life and Thought. New Haven: Yale University Press.Google Scholar
Miao, F. J. P., Green, P. G., Coderre, T. J., Jänig, W. and Levine, J. D. (1996a). Sympathetic-dependence in bradykinin-induced synovial plasma extravasation is dose-related. Neurosci. Lett., 205, 165–168.CrossRefGoogle Scholar
Miao, F. J. P., Jänig, W. and Levine, J. D. (1996b). Role of sympathetic postganglionic neurons in synovial plasma extravasation induced by bradykinin. J. Neurophysiol., 75, 715–724.CrossRefGoogle Scholar
Miao, F. J. P., Jänig, W., Green, P. G. and Levine, J. D. (1997a). Inhibition of bradykinin-induced synovial plasma extravasation produced by noxious cutaneous and visceral stimuli and its modulation by activity in the vagal nerve. J. Neurophysiol., 78, 1285–1292.CrossRefGoogle Scholar
Miao, F. J. P., Jänig, W. and Levine, J. D. (1997b). Vagal branches involved in inhibition of bradykinin-induced synovial plasma extravasation by intrathecal nicotine and noxious stimulation in the rat. J. Physiol., 498, 473–481.CrossRefGoogle Scholar
Miao, F. J. P., Jänig, W. and Levine, J. D. (2000). Nociceptive neuroendocrine negative feedback control of neurogenic inflammation activated by capsaicin in the rat paw: role of the adrenal medulla. J. Physiol., 527, 601–610.CrossRefGoogle ScholarPubMed
Miao, F. J. P., Jänig, W., Jasmin, L. and Levine, J. D. (2001). Spino-bulbospinal pathway mediating vagal modulation of nociceptive-neuroendocrine control of inflammation in the rat. J. Physiol., 532, 811–822.CrossRefGoogle Scholar
Miao, F. J. P., Green, P. G. and Levine, J. D. (2003b). Mechano-sensitive duodenal afferents contribute to vagal modulation of inflammation in the rat. J. Physiol., 554, 227–235.CrossRefGoogle Scholar
Miao, F. J. P., Jänig, W., Jasmin, L. and Levine, J. D. (2003a). Blockade of nociceptive inhibition of plasma extravasation by opioid stimulation of the periaqueductal gray and its interaction with vagus-induced inhibition in the rat. Neuroscience, 119, 875–885.CrossRefGoogle Scholar
Michaelis, M., Göder, R., Häbler, H. J. and Jänig, W. (1994). Properties of afferent nerve fibres supplying the saphenous vein in the cat. J. Physiol., 474, 233–243.CrossRefGoogle ScholarPubMed
Michaelis, M., Häbler, H. J. and Jänig, W. (1996). Silent afferents: a separate class of primary afferents?Clin. Exp. Pharmacol. Physiol., 23, 99–105.CrossRefGoogle ScholarPubMed
Michl, T., Jocic, M., Heinemann, A., Schuligoi, R. and Holzer, P. (2001). Vagal afferent signaling of a gastric mucosal acid insult to medullary, pontine, thalamic, hypothalamic and limbic, but not cortical, nuclei of the rat brain. Pain, 92, 19–27.CrossRefGoogle Scholar
Mifflin, S. W., Spyer, K. M. and Withington-Wray, D. J. (1988). Baroreceptor inputs to the nucleus tractus solitarius in the cat: postsynaptic actions and the influence of respiration. J. Physiol., 399, 349–367.CrossRefGoogle Scholar
Milner, T. A., Morrison, S. F., Abate, C. and Reis, D. J. (1988). Phenylethanolamine N-methyltransferase-containing terminals synapse directly on sympathetic preganglionic neurons in the rat. Brain Res., 448, 205–222.CrossRefGoogle ScholarPubMed
Minson, J. B., Llewellyn-Smith, I. J., Chalmers, J. P., Pilowsky, P. M. and Arnolda, L. F. (1997). c-fos identifies γ-aminobutyric acid-synthesizing barosensitive neurons in caudal ventrolateral medulla. NeuroReport, 8, 3015–3021.CrossRefGoogle Scholar
Mitchell, J. H. (1985). Cardiovascular control during exercise: central and reflex neural mechanisms. Am. J. Cardiol., 55, 34D–41D.CrossRefGoogle ScholarPubMed
Mitchell, R. A., Herbert, D. A., Baker, D. G. and Basbaum, C. B. (1987). In vivo activity of tracheal parasympathetic ganglion cells innervating tracheal smooth muscle. Brain Res., 437, 157–160.CrossRefGoogle ScholarPubMed
Mitchell, S. W. (1872). Injuries of Nerves and their Consequences. Philadelphia: JB Lippincott.Google Scholar
Miura, M., Takayama, K. and Okada, J. (1994). Distribution of glutamate- and γ-aminobutyric acid-immunoreactive neurons projecting to the cardioacceleratory center of the intermediolateral nucleus of the thoracic cord of SHR and WKY rats: a double-labeling study. Brain Res., 638, 139–150.CrossRefGoogle Scholar
Miyawaki, T., Goodchild, A. K. and Pilowsky, P. M. (2003). Maintenance of sympathetic tone by a nickel chloride sensitive mechanism in the rostral ventrolateral medulla of the adult rat. Neuroscience, 116, 455–464.CrossRefGoogle ScholarPubMed
Molander, C. and Grant, G. (1995). Spinal cord cytoarchitecture. In The Rat Nervous System, ed. Paxinos, G.. San Diego: Academic Press, pp. 39–45.Google Scholar
Monnier, A., Alheid, G. F. and McCrimmon, D. R. (2003). Defining ventral medullary respiratory compartments with a glutamate receptor agonist in the rat. J. Physiol., 548, 859–874.CrossRefGoogle ScholarPubMed
Moore, R. Y. (1996). Neural control of the pineal gland. Behav. Brain Res., 73, 125–130.CrossRefGoogle ScholarPubMed
Moore, R. Y. (2003). Circadian timing. In Fundamental Neuroscience, 2nd edn., ed. Squire, L. R., Bloom, F. E., McConnell, S. K.et al. San Diego: Academic Press, pp. 1067–1084.Google Scholar
Moran, T. H. and Schwartz, G. J. (1994). Neurobiology of cholecystokinin. Crit. Rev. Neurobiol., 9, 1–28.Google ScholarPubMed
Morgan, C. W., Groat, W. C., Felkins, L. A. and Zhang, S. J. (1993). Intracellular injection of neurobiotin or horseradish peroxidase reveals separate types of preganglionic neurons in the sacral parasympathetic nucleus of the cat. J. Comp. Neurol., 331, 161–182.CrossRefGoogle ScholarPubMed
Moriarty, M., Gibbins, I. L., Potter, E. K. and McCloskey, D. I. (1992). Comparison of the inhibitory roles of neuropeptide Y and galanin on cardiac vagal action in the dog. Neurosci. Lett., 139, 275–279.CrossRefGoogle ScholarPubMed
Morris, J. L. (1995). Distribution and peptide content of sympathetic axons innervating different regions of the cutaneous venous bed in the pinna of the guinea pig ear. J. Vasc. Res., 32, 378–386.CrossRefGoogle ScholarPubMed
Morris, J. L. (1999). Cotransmission from sympathetic vasoconstrictor neurons to small cutaneous arteries in vivo. Am. J. Physiol., 277, H58–H64.Google ScholarPubMed
Morris, J. and Dolan, R. (2004). Functional neuroanatomy of human emotion. In Human Brain Function, 2nd edn., ed. Frackowiak, R. S. J., Friston, K. J., Frith, C. D.et al. Amsterdam: Elsevier Academic Press, pp. 365–396.Google Scholar
Morris, J. L. and Gibbins, I. L. (1992). Co-transmission and neuromodulation. In The Autonomic Nervous System, Vol. 1, Autonomic Neuroeffector Mechanisms, ed. Burnstock, G. and Hoyle, C. H. V.. Chur, Switzerland: Harwood, pp. 33–119.Google Scholar
Morris, J. L. and Gibbins, I. L. (eds.) (1997). The Autonomic Nervous System, Vol. 12, Autonomic Innervation of the Skin, Chur, Switzerland: Harwood Academic Publishers.Google Scholar
Morris, J. L., Gibbins, I. L. and Clevers, J. (1981). Resistance of adrenergic neurotransmission in the toad heart to adrenoceptor blockade. Naunyn-Schmiedeberg's Arch. Pharmacol., 317, 331–338.CrossRefGoogle ScholarPubMed
Morris, M. J., Russell, A. E., Kapoor, V.et al. (1986). Increases in plasma neuropeptide Y concentrations during sympathetic activation in man. J. Auton. Nerv. Syst., 17, 143–149.CrossRefGoogle ScholarPubMed
Morris, J. L., Gibbins, I. L. and Jobling, P. (2005). Post-stimulus potentiation of transmission in pelvic ganglia enhances sympathetic dilatation of guinea-pig uterine artery in vitro. J. Physiol., 566, 189–203.CrossRefGoogle ScholarPubMed
Morrison, J. F. B. (1997). Central nervous control of the bladder. In The Autonomic Nervous System, Vol. 11, Central Nervous Control of Autonomic Function, ed. Jordan, D.. Amsterdam: Harwood Academic Publishers, pp. 129–149.Google Scholar
Morrison, S. F. (1999). rostral ventrolateral medulla and raphe differentially regulate sympathetic outflows to splanchnic and brown adipose tissue. Am. J. Physiol., 276, R962–R973.Google ScholarPubMed
Morrison, S. F. (2001a). Differential control of sympathetic outflow. Am. J. Physiol. Regul. Integr. Comp. Physiol., 281, R683–R698.CrossRefGoogle Scholar
Morrison, S. F. (2001b). Differential regulation of brown adipose and splanchnic sympathetic outflows in rat: roles of raphe and rostral ventrolateral medulla neurons. Clin. Exp. Pharmacol. Physiol., 28, 138–143.CrossRefGoogle Scholar
Morrison, S. F. and Cao, W. H. (2000). Different adrenal sympathetic preganglionic neurons regulate epinephrine and norepinephrine secretion. Am. J. Physiol. Regul. Integr. Comp. Physiol., 279, R1763–R1775.CrossRefGoogle ScholarPubMed
Morrison, S. F. and Gebber, G. L. (1985). Axonal branching patterns and funicular trajectories of raphespinal sympathoinhibitory neurons. J. Neurophysiol., 53, 759–772.CrossRefGoogle ScholarPubMed
Morrison, S. F., Callaway, J., Milner, T. A. and Reis, D. J. (1991). Rostral ventrolateral medulla: a source of the glutamatergic innervation of the sympathetic intermediolateral nucleus. Brain Res., 562, 126–135.CrossRefGoogle ScholarPubMed
Morrison, S. F., Sved, A. F. and Passerin, A. M. (1999). γ-aminobutyric acid-mediated inhibition of raphe pallidus neurons regulates sympathetic outflow to brown adipose tissue. Am. J. Physiol., 276, R290–R297.Google Scholar
Mosqueda-Garcia, R., Furlan, R., Tank, J. and Fernandez-Violante, R. (2000). The elusive pathophysiology of neurally mediated syncope. Circulation, 102, 2898–2906.CrossRefGoogle ScholarPubMed
Müller, L. R. (1902). Klinische und experimentelle Studien ueber die Innervation der Blase, des Mastdarmes und des Genitalapparates [Clinical and experimental studies on the innervation of the urinary bladder, hindgut and reproductive organs]. Dtsch. Z. Nervenheilkd., 21, 86–155.CrossRefGoogle Scholar
Müller, L. R. (1906). Ueber die Exstirpation der unteren Haelfte des Rueckenmarkes und deren Folgeerscheinungen [On the consequences of the removal of the lower half of the spinal cord]. Dtsch. Z. Nervenheilkd., 30, 411–423.CrossRefGoogle Scholar
Mulryan, K., Gitterman, D. P., Lewis, C. J.et al. (2000). Reduced vas deferens contraction and male infertility in mice lacking P2X1 receptors. Nature, 403, 86–89.CrossRefGoogle ScholarPubMed
Mulvany, M. J., Nilsson, H. and Flatman, J. A. (1982). Role of membrane potential in the response of rat small mesenteric arteries to exogenous noradrenaline stimulation. J. Physiol., 332, 363–373.CrossRefGoogle ScholarPubMed
Murphy, S. M., Matthew, S. E., Rodgers, H. F., Lituri, D. T. and Gibbins, I. L. (1998). Synaptic organisation of neuropeptide-containing preganglionic boutons in lumbar sympathetic ganglia of guinea pigs. J. Comp. Neurol., 398, 551–567.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Murray, J. G. and Thompson, J. W. (1957). The occurrence and function of collateral sprouting in the sympathetic nervous system of the cat. J. Physiol., 135, 133–162.CrossRefGoogle ScholarPubMed
Nadelhaft, I. and Vera, P. L. (1995). Central nervous system neurons infected by pseudorabies virus injected into the rat urinary bladder following unilateral transection of the pelvic nerve. J. Comp. Neurol., 359, 443–456.CrossRefGoogle ScholarPubMed
Nadelhaft, I., Vera, P. L., Card, J. P. and Miselis, R. R. (1992). Central nervous system neurons labelled following the injection of pseudorabies virus into the rat urinary bladder. Neurosci. Lett., 143, 271–274.CrossRefGoogle ScholarPubMed
Nagashima, K., Nakai, S., Tanaka, M. and Kanosue, K. (2000). Neuronal circuitries involved in thermoregulation. Auton. Neurosci., 85, 18–25.CrossRefGoogle ScholarPubMed
Natarajan, M. and Morrison, S. F. (2000). Sympathoexcitatory caudal ventrolateral medulla neurons mediate responses to caudal pressor area stimulation. Am. J. Physiol. Regul. Integr. Comp. Physiol., 279, R364–R374.CrossRefGoogle ScholarPubMed
Neff, R. A., Mihalevich, M. and Mendelowitz, D. (1998). Stimulation of nucleus tractus solitarii activates N-methyl-D-aspartate (acid) and non-N-methyl-D-aspartate (acid) receptors in rat cardiac vagal neurons in the nucleus ambiguus. Brain Res., 792, 277–282.CrossRefGoogle Scholar
Ness, T. J. and Gebhart, G. F. (1990). Visceral pain: a review of experimental studies. Pain, 41, 167–234.CrossRefGoogle ScholarPubMed
Neuhuber, W. L. (1989). Vagal afferent fibers almost exclusively innervate islets in the rat pancreas as demonstrated by anterograde tracing. J. Auton. Nerv. Syst., 29, 13–18.CrossRefGoogle ScholarPubMed
Niijima, A. (1996). The afferent dischargee from sensors for interleukin 1 beta in the hepatoportal system in the anesthetized rat. J. Auton. Nerv. Syst., 61, 287–291.CrossRefGoogle Scholar
Nilsson, H., Jensen, P. E. and Mulvany, M. J. (1994). Minor role for direct adrenoceptor-mediated calcium entry in rat mesenteric small arteries. J. Vasc. Res., 31, 314–321.CrossRefGoogle ScholarPubMed
Nilsson, S. (1983). Autonomic Nerve Function in the Vertebrates. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Nilsson, S. and Holmgren, S. (eds.) (1994). The Autonomic Nervous System, Vol. 4, Comparative Physiology and Evolution of the Autonomic Nervous System, Chur, Switzerland: Harwood Academic Publishers.Google Scholar
Nisida, I. and Okada, H. (1960). The activity of the pupilloconstrictor centers. Jap. J. Physiol., 10, 64–72.CrossRefGoogle Scholar
Nja, A. and Purves, D. (1977a). Specific innervation of guinea-pig superior cervical ganglion cells by preganglionic fibres arising from different levels of the spinal cord. J. Physiol., 264, 565–583.CrossRefGoogle Scholar
Nja, A. and Purves, D. (1977b). Re-innervation of guinea-pig superior cervical ganglion cells by preganglionic fibres arising from different levels of the spinal cord. J. Physiol., 272, 633–651.CrossRefGoogle Scholar
Nolan, M. F., Logan, S. D. and Spanswick, D. (1999). Electrophysiological properties of electrical synapses between rat sympathetic preganglionic neurones in vitro. J. Physiol., 519, 753–764.CrossRefGoogle ScholarPubMed
Noll, G., Elam, M., Kunimoto, M., Karlsson, T. and Wallin, B. G. (1994). Skin sympathetic nerve activity and effector function during sleep in humans. Acta Physiol. Scand., 151, 319–329.CrossRefGoogle ScholarPubMed
Nordin, M. (1990). Sympathetic discharges in the human supraorbital nerve and their relation to sudo- and vasomotor responses. J. Physiol., 423, 241–255.CrossRefGoogle ScholarPubMed
Nordin, M. and Fagius, J. (1995). Effect of noxious stimulation on sympathetic vasoconstrictor outflow to human muscles. J. Physiol., 489, 885–894.CrossRefGoogle ScholarPubMed
Norman, R. A., Coleman, T. G.Jr and Dent, A. C. (1981). Continuous monitoring of arterial pressure indicates sinoaortic denervated rats are not hypertensive. Hypertension, 3, 119–125.CrossRefGoogle Scholar
North, R. A. (1986). Receptors on individual neurones. Neuroscience, 17, 899–907.CrossRefGoogle ScholarPubMed
Oberle, J., Elam, M., Karlsson, T. and Wallin, B. G. (1988). Temperature-dependent interaction between vasoconstrictor and vasodilator mechanisms in human skin. Acta Physiol. Scand., 132, 459–469.CrossRefGoogle ScholarPubMed
Obrist, P. A. (1981). Cardiovascular Psychophysiology: A Perspective. New York: Plenum Press.CrossRefGoogle Scholar
Ochoa, J. and Torebjörk, E. (1983). Sensations evoked by intraneural microstimulation of single mechanoreceptor units innervating the human hand. J. Physiol., 342, 633–654.CrossRefGoogle ScholarPubMed
Ochoa, J. and Torebjörk, H. E. (1989). Sensations evoked by intraneural microstimulation of C nociceptor fibres in human skin nerves. J. Physiol., 415, 583–599.CrossRefGoogle ScholarPubMed
Oh, E. J., Mazzone, S. B., Canning, B. J. and Weinreich, D. (2006). Reflex regulation of airway sympathetic nerves in guinea-pigs. J. Physiol., in press.CrossRefGoogle ScholarPubMed
Oldfield, B. J. and McLachlan, E. M. (1981). An analysis of the sympathetic preganglionic neurons projecting from the upper thoracic spinal roots of the cat. J. Comp. Neurol., 196, 329–345.CrossRefGoogle ScholarPubMed
Öngür, D. and Price, J. L. (2000). The organization of networks within the orbital and medial prefrontal cortex of rats, monkeys and humans. Cereb. Cortex, 10, 206–219.CrossRefGoogle ScholarPubMed
Öngür, D., An, X. and Price, J. L. (1998). Prefrontal cortical projections to the hypothalamus in macaque monkeys. J. Comp. Neurol., 401, 480–505.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Ootsuka, Y. and Terui, N. (1997). Functionally different neurons are organized topographically in the rostral ventrolateral medulla of rabbits. J. Auton. Nerv. Syst., 67, 67–78.CrossRefGoogle ScholarPubMed
Orr, R. and Marson, L. (1998). Identification of central nervous system neurons innervating the rat prostate: a transneuronal tracing study using pseudorabies virus. J. Auton. Nerv. Syst., 72, 4–15.CrossRefGoogle ScholarPubMed
Owens, N. C., Ootsuka, Y., Kanosue, K. and McAllen, R. M. (2002). Thermoregulatory control of sympathetic fibers supplying the rat's tail. J. Physiol., 543, 849–858.CrossRefGoogle ScholarPubMed
Owsjannikow, P. (1874). Über einen Unterschied in den reflectorischen Leistungen des verlängerten und des Rückenmarkes der Kaninchen [On a difference between reflexes mediated by the medulla oblongata and spinal cord in rabbits]. Sitzungsber. Akad. Wiss. Wien, Math.-Naturw., Abt. 2, 26, 457–464.Google Scholar
Ozaki, N. and Gebhart, G. F. (2001). Characterization of mechanosensitive splanchnic nerve afferent fibers innervating the rat stomach. Am. J. Physiol. Gastrointest. Liver Physiol., 281, G1449–G1459.CrossRefGoogle ScholarPubMed
Page, A. J. and Blackshaw, L. A. (1998). An in vitro study of the properties of vagal afferent fibres innervating the ferret oesophagus and stomach. J. Physiol., 512, 907–916.CrossRefGoogle Scholar
Page, A. J., Martin, C. M. and Blackshaw, L. A. (2002). Vagal mechanoreceptors and chemoreceptors in mouse stomach and esophagus. J. Neurophysiol., 87, 2095–2103.CrossRefGoogle ScholarPubMed
Paintal, A. S. (1973). Vagal sensory receptors and their reflex effects. Physiol. Rev., 53, 159–227.CrossRefGoogle ScholarPubMed
Paintal, A. S. (1986). The visceral sensations – some basic mechanisms. Prog. Brain Res., 67, 3–19.CrossRefGoogle ScholarPubMed
Pan, H. L. and Longhurst, J. C. (1996). Ischaemia-sensitive sympathetic afferents innervating the gastrointestinal tract function as nociceptors in cats. J. Physiol., 492, 841–850.CrossRefGoogle ScholarPubMed
Pan, H. L., Longhurst, J. C., Eisenach, J. C. and Chen, S. R. (1999). Role of protons in activation of cardiac sympathetic C-fibre afferents during ischaemia in cats. J. Physiol., 518, 857–866.CrossRefGoogle ScholarPubMed
Panksepp, J. (1998). Affective Neuroscience. New York, Oxford: Oxford University Press.Google Scholar
Papka, R. E., Williams, S., Miller, K. E., Copelin, T. and Puri, P. (1998). central nervous system location of uterine-related neurons revealed by trans-synaptic tracing with pseudorabies virus and their relation to estrogen receptor-immunoreactive neurons. Neuroscience, 84, 935–952.CrossRefGoogle ScholarPubMed
Parr, E. J. and Sharkey, K. A. (1996). Immunohistochemically-defined subtypes of neurons in the inferior mesenteric ganglion of the guinea-pig. J. Auton. Nerv. Syst., 59, 140–150.CrossRefGoogle ScholarPubMed
Paton, J. F. (1996a). The ventral medullary respiratory network of the mature mouse studied in a working heart–brainstem preparation. J. Physiol., 493, 819–831.CrossRefGoogle Scholar
Paton, J. F. (1996b). A working heart–brainstem preparation of the mouse. J. Neurosci. Methods, 65, 63–68.CrossRefGoogle Scholar
Paton, J. F. (1999). The Sharpey-Schafer prize lecture: nucleus tractus solitarii: integrating structures. Exp. Physiol., 84, 815–833.CrossRefGoogle ScholarPubMed
Paton, J. F. and Dutschmann, M. (2002). Central control of upper airway resistance regulating respiratory airflow in mammals. J. Anat., 201, 319–323.CrossRefGoogle ScholarPubMed
Paton, J. F. and Kasparov, S. (2000). Sensory channel specific modulation in the nucleus of the solitary tract. J. Auton. Nerv. Syst., 80, 117–129.CrossRefGoogle ScholarPubMed
Paton, J. F., Li, Y. W., Deuchars, J. and Kasparov, S. (2000). Properties of solitary tract neurons receiving inputs from the sub-diaphragmatic vagus nerve. Neuroscience, 95, 141–153.CrossRefGoogle ScholarPubMed
Paton, J. F., Deuchars, J., Li, Y. W. and Kasparov, S. (2001). Properties of solitary tract neurones responding to peripheral arterial chemoreceptors. Neuroscience, 105, 231–248.CrossRefGoogle ScholarPubMed
Paton, J. F., Deuchar, J., Wang, S. and Kasparov, S. (2005). Nitroxergic modulation in the nucleus tractus solitarii: implications for cardiovascular function. In Advances in Vagal Afferent Neurobiology, ed. Undem, B. and Weinreich, D.. Boca Raton: CRC Press, pp. 209–246.CrossRefGoogle Scholar
Paxinos, G. and Watson, C. (eds.) (1998). The Rat Brain. San Diego: Academic Press.Google Scholar
Peng, H., Matchkov, V., Ivarsen, A., Aalkjaer, C. and Nilsson, H. (2001). Hypothesis for the initiation of vasomotion. Circ. Res., 88, 810–815.CrossRefGoogle ScholarPubMed
Pernow, J. (1988). Co-release and functional interactions of neuropeptide Y and noradrenaline in peripheral sympathetic vascular control. Acta Physiol. Scand. Suppl., 568, 1–56.Google ScholarPubMed
Perry, M. J. and Lawson, S. N. (1998). Differences in expression of oligosaccharides, neuropeptides, carbonic anhydrase and neurofilament in rat primary afferent neurons retrogradely labelled via skin, muscle or visceral nerves. Neuroscience, 85, 293–310.CrossRefGoogle ScholarPubMed
Persson, P., Ehmke, H., Kirchheim, H. and Seller, H. (1988). Effect of sino-aortic denervation in comparison to cardiopulmonary deafferentiation on long-term blood pressure in conscious dogs. Pflügers Arch., 411, 160–166.CrossRefGoogle ScholarPubMed
Petras, J. M. and Cummings, J. F. (1972). Autonomic neurons in the spinal cord of the Rhesus monkey: a correlation of the findings of cytoarchitectonics and sympathectomy with fiber degeneration following dorsal rhizotomy. J. Comp. Neurol., 146, 189–218.CrossRefGoogle ScholarPubMed
Petrovich, G. D., Canteras, N. S. and Swanson, L. W. (2001). Combinatorial amygdalar inputs to hippocampal domains and hypothalamic behavior systems. Brain Res. Brain Res. Rev., 38, 247–289.CrossRefGoogle ScholarPubMed
Petty, B. G., Cornblath, D. R., Adornato, B. T.et al. (1994). The effect of systemically administered recombinant human nerve growth factor in healthy human subjects. Ann. Neurol., 36, 244–246.CrossRefGoogle ScholarPubMed
Phan, K. L., Wager, T., Taylor, S. F. and Liberzon, I. (2002). Functional neuroanatomy of emotion: a meta-analysis of emotion activation studies in PET and fMRI. Neuroimage, 16, 331–348.CrossRefGoogle ScholarPubMed
Phillips, J. K., Goodchild, A. K., Dubey, R.et al. (2001). Differential expression of catecholamine biosynthetic enzymes in the rat ventrolateral medulla. J. Comp. Neurol., 432, 20–34.CrossRefGoogle ScholarPubMed
Phillips, R. J. and Powley, T. L. (2000). Tension and stretch receptors in gastrointestinal smooth muscle: re-evaluating vagal mechanoreceptor electrophysiology. Brain Res. Brain Res. Rev., 34, 1–26.CrossRefGoogle ScholarPubMed
Phillips, R. J., Baronowsky, E. A. and Powley, T. L. (1997). Afferent innervation of gastrointestinal tract smooth muscle by the hepatic branch of the vagus. J. Comp. Neurol., 384, 248–270.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Pick, J. (1970). The Autonomic Nervous System. Philadelphia: Lippincott.Google Scholar
Pickering, A. E., Boscan, P. and Paton, J. F. (2003). Nociception attenuates parasympathetic but not sympathetic baroreflex via neurokinin 1 receptors in the rat nucleus tractus solitarii. J. Physiol., 551, 589–599.CrossRefGoogle Scholar
Pierce, P. A., Xie, G. X., Peroutka, S. J., Green, P. G. and Levine, J. D. (1995). 5-Hydroxytryptamine-induced synovial plasma extravasation is mediated via 5-hydroxytryptamine2 A receptors on sympathetic efferent terminals. J. Pharmacol. Exp. Ther., 275, 502–508.Google Scholar
Pilowsky, P. M. and Goodchild, A. K. (2002). Baroreceptor reflex pathways and neurotransmitters: 10 years on. J. Hypertens., 20, 1675–1688.CrossRefGoogle Scholar
Pilowsky, P. M. and Makeham, J. (2001). Juxtacellular labeling of identified neurons: kiss the cells and make them dye. J. Comp. Neurol., 433, 1–3.CrossRefGoogle ScholarPubMed
Pilowsky, P. M., Jiang, C. and Lipski, J. (1990). An intracellular study of respiratory neurons in the rostral ventrolateral medulla of the rat and their relationship to catecholamine-containing neurons. J. Comp. Neurol., 301, 604–617.CrossRefGoogle ScholarPubMed
Pilowsky, P., Llewellyn-Smith, I. J., Arnolda, L., Minson, J. and Chalmers, J. (1994). Intracellular recording from sympathetic preganglionic neurons in cat lumbar spinal cord. Brain Res., 656, 319–328.CrossRefGoogle ScholarPubMed
Pinault, D. (1996). A novel single-cell staining procedure performed in vivo under electrophysiological control: morpho-functional features of juxtacellularly labeled thalamic cells and other central neurons with biocytin or Neurobiotin. J. Neurosci. Methods, 65, 113–136.CrossRefGoogle ScholarPubMed
Polgar, E., Gray, S., Riddell, J. S. and Todd, A. J. (2004). Lack of evidence for significant neuronal loss in laminae I-III of the spinal dorsal horn of the rat in the chronic constriction injury model. Pain, 111, 144–150.CrossRefGoogle ScholarPubMed
Polson, J. W., Halliday, G. M., McAllen, R. M., Coleman, M. J. and Dampney, R. A. (1992). Rostrocaudal differences in morphology and neurotransmitter content of cells in the subretrofacial vasomotor nucleus. J. Auton. Nerv. Syst., 38, 117–138.CrossRefGoogle ScholarPubMed
Polson, J. W., Potts, P. D., Li, Y. W. and Dampney, R. A. (1995). Fos expression in neurons projecting to the pressor region in the rostral ventrolateral medulla after sustained hypertension in conscious rabbits. Neuroscience, 67, 107–123.CrossRefGoogle ScholarPubMed
Popescu, L. M., Ciontea, S. M., Cretoiu, D.et al. (2005a). Novel type of interstitial cell (Cajal-like) in human fallopian tube. J. Cell Mol. Med., 9, 479–523.CrossRefGoogle Scholar
Popescu, L. M., Hinescu, M. E., Ionescu, N.et al. (2005b). Interstitial cells of Cajal in pancreas. J. Cell Mol. Med., 9, 169–190.CrossRefGoogle Scholar
Possas, O. S., Campos, R. R. Jr., Cravo, S. L., Lopes, O. U. and Guertzenstein, P. G. (1994). A fall in arterial blood pressure produced by inhibition of the caudalmost ventrolateral medulla: the caudal pressor area. J. Auton. Nerv. Syst., 49, 235–245.CrossRefGoogle ScholarPubMed
Potter, E. K. (1987). Guanethidine blocks neuropeptide-Y-like inhibitory action of sympathetic nerves on cardiac vagus. J. Auton. Nerv. Syst., 21, 87–90.CrossRefGoogle ScholarPubMed
Potter, E. K. (1991). Neuropeptide Y as an autonomic neurotransmitter. In Novel Peripheral Neurotransmitter, ed. Bell, C.. New York: Pergamon, pp. 81–112.Google Scholar
Powley, T. L., Berthoud, H. R., Fox, A. P. and Laughton, W. (1992). The dorsal vagal complex forms a sensory-motor lattice: the circuitry of gastrointestinal reflexes. In Neuroanatomy and Physiology of Abdominal Vagal Afferents, ed. Ritter, S., Ritter, R. C. and Barnes, C. D.. Boca Raton: CRC Press, pp. 55–79.Google Scholar
Prechtl, J. C. and Powley, T. L. (1985). Organization and distribution of the rat subdiaphragmatic vagus and associated paraganglia. J. Comp. Neurol., 235, 182–195.CrossRefGoogle ScholarPubMed
Prechtl, J. C. and Powley, T. L. (1990a). The fiber composition of the abdominal vagus of the rat. Anat. Embryol. (Berl.), 181, 101–115.CrossRefGoogle Scholar
Prechtl, J. C. and Powley, T. L. (1990b). B-afferents: a fundamental division of the nervous system mediating homeostasis [the article includes Open Peer Commentary]. Behav. Brain Sci., 13, 289–331.CrossRefGoogle Scholar
Purves, D. (1988). Body and Brain: A Tropic Theory of Neural Connection. Cambridge: Harvard University Press.Google Scholar
Purves, D. and Hume, R. I. (1981). The relation of postsynaptic geometry to the number of presynaptic axons that innervate autonomic ganglion cells. J. Neurosci., 1, 441–452.CrossRefGoogle ScholarPubMed
Purves, D. and Lichtman, J. W. (1978). Formation and maintenance of synaptic connections in autonomic ganglia. Physiol. Rev., 58, 821–862.CrossRefGoogle ScholarPubMed
Purves, D. and Lichtman, J. W. (1985). Geometrical differences among homologous neurons in mammals. Science, 228, 298–302.CrossRefGoogle ScholarPubMed
Purves, D. and Wigston, D. J. (1983). Neural units in the superior cervical ganglion of the guinea-pig. J. Physiol., 334, 169–178.CrossRefGoogle ScholarPubMed
Purves, D., Rubin, E., Snider, W. D. and Lichtman, J. (1986). Relation of animal size to convergence, divergence, and neuronal number in peripheral sympathetic pathways. J. Neurosci., 6, 158–163.CrossRefGoogle ScholarPubMed
Purves, D., Snider, W. D. and Voyvodic, J. T. (1988). Trophic regulation of nerve cell morphology and innervation in the autonomic nervous system. Nature, 336, 123–128.CrossRefGoogle ScholarPubMed
Pyner, S. and Coote, J. H. (1994). Evidence that sympathetic preganglionic neurons are arranged in target-specific columns in the thoracic spinal cord of the rat. J. Comp. Neurol., 342, 15–22.CrossRefGoogle ScholarPubMed
Pyner, S. and Coote, J. H. (1998). Rostroventrolateral medulla neurons preferentially project to target-specified sympathetic preganglionic neurons. Neuroscience, 83, 617–631.CrossRefGoogle ScholarPubMed
Qin, C., Chandler, M. J., Miller, K. E. and Foreman, R. D. (2001). Responses and afferent pathways of superficial and deeper c(1)–c(2) spinal cells to intrapericardial algogenic chemicals in rats. J. Neurophysiol., 85, 1522–1532.CrossRefGoogle ScholarPubMed
Quigg, M., Elfvin, L. G. and Aldskogius, H. (1990). Anterograde transsynaptic transport of WGA-horse radish peroxidase from spinal afferents to postganglionic sympathetic cells of the stellate ganglion of the guinea pig. Brain Res., 518, 173–178.CrossRefGoogle Scholar
Randall, W. C. (ed.) (1984). Nervous Control of Cardiovascular Function. New York, Oxford: Oxford University Press.Google Scholar
Randall, W. C., Wurster, R. D. and Lewin, R. J. (1966). Responses of patients with high spinal transection to high ambient temperatures. J. Appl. Physiol., 21, 985–993.CrossRefGoogle ScholarPubMed
Randich, A. and Gebhart, G. F. (1992). Vagal afferent modulation of nociception. Brain Res. Rev., 17, 77–99.CrossRefGoogle ScholarPubMed
Ranson, S. W. and Clark, S. L. (1959). The Anatomy of the Nervous System. 10th edn. Philadelphia, London: W. B. Saunders Company.Google Scholar
Ranson, S. W. and Magoun, H. W. (1939). The hypothalamus. Ergebn. Physiol., 41, 56–163.CrossRefGoogle Scholar
Rathner, J. A. and McAllen, R. M. (1999). Differential control of sympathetic drive to the rat tail artery and kidney by medullary premotor cell groups. Brain Res., 834, 196–199.CrossRefGoogle ScholarPubMed
Rathner, J. A., Owens, N. C. and McAllen, R. M. (2001). Cold-activated raphe-spinal neurons in rats. J. Physiol., 535, 841–854.CrossRefGoogle ScholarPubMed
Reed, D. E. and Vanner, S. J. (2003). Long vasodilator reflexes projecting through the myenteric plexus in guinea-pig ileum. J. Physiol., 553, 911–924.CrossRefGoogle ScholarPubMed
Reiner, A., Karten, H. J., Gamlin, P. D. R. and Erichsen, J. T. (1983). Parasympathetic ocular control. Functional subdivisions and circuitry of the avian nucleus Edinger–Westphal. Trends Neurosci., 6, 140–145.CrossRefGoogle Scholar
Reis, D. J., Golanov, E. V., Ruggiero, D. A. and Sun, M. K. (1994). Sympatho-excitatory neurons of the rostral ventrolateral medulla are oxygen sensors and essential elements in the tonic and reflex control of the systemic and cerebral circulations. J. Hypertens. Suppl., 12, S159–S180.Google ScholarPubMed
Reis, D. J., Golanov, E. V., Galea, E. and Feinstein, D. L. (1997). Central neurogenic neuroprotection: central neural systems that protect the brain from hypoxia and ischemia. Ann. New York Acad. Sci., 835, 168–186.CrossRefGoogle ScholarPubMed
Rekling, J. C. and Feldman, J. L. (1998). PreBotzinger complex and pacemaker neurons: hypothesized site and kernel for respiratory rhythm generation. Annu. Rev. Physiol., 60, 385–405.CrossRefGoogle ScholarPubMed
Rexed, B. (1952). The cytoarchitectonic organization of the spinal cord in the cat. J. Comp. Neurol., 96, 414–495.CrossRefGoogle ScholarPubMed
Rexed, B. (1954). A cytoarchitectonic atlas of the spinal cord in the cat. J. Comp. Neurol., 100, 297–379.CrossRefGoogle ScholarPubMed
Richards, W., Hillsley, K., Eastwood, C. and Grundy, D. (1996). Sensitivity of vagal mucosal afferents to cholecystokinin and its role in afferent signal transduction in the rat. J. Physiol., 497, 473–481.CrossRefGoogle ScholarPubMed
Richter, D. W. (1982). Generation and maintenance of the respiratory rhythm. J. Exp. Biol., 100, 93–107.Google ScholarPubMed
Richter, D. W. (1996). Neural regulation of respiration: rhythmogenesis and afferent control. In Comprehensive Human Physiology, ed. Greger, R. and Windhorst, U.. Berlin, Heidelberg, New York: Springer, pp. 2079–2095.CrossRefGoogle Scholar
Richter, D. W. and Ballantyne, D. (1983). A three phase theory about the basic respiratory pattern generator. In Central Neurone Environment, ed. Schläfke, M. E., Koepchen, H. P. and See, W. R.. Berlin, Heidelberg, New York: Springer, pp. 164–174.Google Scholar
Richter, D. W. and Spyer, K. M. (1990). Cardiorespiratory control. In Central Regulation of Autonomic Functions, eds. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 189–207.Google Scholar
Richter, D. W. and Spyer, K. M. (2001). Studying rhythmogenesis of breathing: comparison of in vivo and in vitro models. Trends Neurosci., 24, 464–472.CrossRefGoogle ScholarPubMed
Richter, D. W., Spyer, K. M., Gilbey, M. P. et al. (1991). On the existence of a common cardiorespiratory network. In Cardiorespiratory and Motor Coordination, ed. Koepchen, H. P. and Huopaniemi, T.. Berlin, Heidelberg, New York: Springer, pp. 118–130.CrossRefGoogle Scholar
Rinaman, L., Card, J. P., Schwaber, J. S. and Miselis, R. R. (1989). Ultrastructural demonstration of a gastric monosynaptic vagal circuit in the nucleus of the solitary tract in rat. J. Neurosci., 9, 1985–1996.CrossRefGoogle ScholarPubMed
Risold, P. Y., Thompson, R. H. and Swanson, L. W. (1997). The structural organization of connections between hypothalamus and cerebral cortex. Brain Res. Brain Res. Rev., 24, 197–254.CrossRefGoogle ScholarPubMed
Ritter, S., Ritter, R. C. and Barnes, C. D. (eds.) (1992). Neuroanatomy and Physiology of Abdominal Vagal Afferents. Boca Raton: CRC Press.Google Scholar
Robertson, D. and Biaggioni, I. (eds.) (1995). The Autonomic Nervous System, Vol. 5, Disorders of the Autonomic Nervous System. Luxembourg: Harwood Academic Publishers.
Roddie, I. C. (1977). Human responses to emotional stress. Irish J. Med. Sci., 146, 395–417.CrossRefGoogle ScholarPubMed
Roddie, I. C., Shepherd, J. T. and Whelan, R. F. (1957). The contribution of constrictor and dilator nerves to the skin vasodilation during body heating. J. Physiol., 136, 489–497.CrossRefGoogle ScholarPubMed
Rogers, R. C. and Hermann, G. E. (1992). Central regulation of brainstem gastric vago-vagal control circuits. In Neuroanatomy and Physiology of Abdominal Vagal Afferents, ed. Ritter, S., Ritter, R. C. and Barnes, C. D.. Boca Raton: CRC Press, pp. 99–134.Google Scholar
Rogers, R. C., Kita, H., Butcher, L. L. and Novin, D. (1980). Afferent projections to the dorsal motor nucleus of the vagus. Brain Res. Bull., 5, 365–373.CrossRefGoogle ScholarPubMed
Rogers, R. C., McTigue, D. M. and Hermann, G. E. (1995). Vagovagal reflex control of digestion: afferent modulation by neural and “endoneurocrine” factors. Am. J. Physiol., 268, G1–G10.Google ScholarPubMed
Rogers, R. C., Hermann, G. E. and Travagli, R. A. (1999). Brainstem pathways responsible for oesophageal control of gastric motility and tone in the rat. J. Physiol., 514, 369–383.CrossRefGoogle ScholarPubMed
Rogers, R. C., Travagli, R. A. and Hermann, G. E. (2003). Noradrenergic neurons in the rat solitary nucleus participate in the esophageal-gastric relaxation reflex. Am. J. Physiol. Regul. Integr. Comp. Physiol., 285, R479–R489.CrossRefGoogle ScholarPubMed
Rogers, R. F., Paton, J. F. and Schwaber, J. S. (1993). nucleus tractus solitarii neuronal responses to arterial pressure and pressure changes in the rat. Am. J. Physiol., 265, R1355–R1368.Google ScholarPubMed
Roman, C. and Gonella, J. (1994). Extrinsic control of digestive tract motility. In Physiology of the Gastrointestinal Tract, 3rd edn., ed. , L. R. Johnson. New York: Raven Press, pp. 507–553.Google Scholar
Romano, T. A., Felten, S. Y., Felten, D. L. and Olschowka, J. A. (1991). Neuropeptide-Y innervation of the rat spleen: another potential immunomodulatory neuropeptide. Brain Behav. Immun., 5, 116–131.CrossRefGoogle ScholarPubMed
Root, W. S. and Bard, P. (1947). The mediation of feline erection through sympathetic pathways with some remarks on sexual behavior after deafferentation of the genitalia. Am. J. Physiol., 151, 80–90.Google ScholarPubMed
Rosell, S. (1980). Neuronal control of microvessels. Annu. Rev. Physiol., 42, 359–371.CrossRefGoogle ScholarPubMed
Rosell, S. and Belfrage, E. (1979). Blood circulation in adipose tissue. Physiol. Rev., 59, 1078–1104.CrossRefGoogle ScholarPubMed
Rosicka, M., Krsek, M., Jarkovska, Z., Marek, J. and Schreiber, V. (2002). Ghrelin – a new endogenous growth hormone secretagogue. Physiol. Res., 51, 435–441.Google ScholarPubMed
Ross, C. A., Ruggiero, D. A., Park, D. H.et al. (1984b). Tonic vasomotor control by the rostral ventrolateral medulla: effect of electrical or chemical stimulation of the area containing C1 adrenaline neurons on arterial pressure, heart rate, and plasma catecholamines and vasopressin. J. Neurosci., 4, 474–494.CrossRefGoogle Scholar
Ross, C. A., Ruggiero, D. A., Joh, T. H., Park, D. H. and Reis, D. J. (1984a). Rostral ventrolateral medulla: selective projections to the thoracic autonomic cell column from the region containing C1 adrenaline neurons. J. Comp. Neurol., 228, 168–185.CrossRefGoogle Scholar
Rossiter, C. D., Norman, W. P., Jain, M.et al. (1990). Control of lower esophageal sphincter pressure by two sites in dorsal motor nucleus of the vagus. Am. J. Physiol., 259, G899–G906.Google ScholarPubMed
Rowell, L. B. (1993). Human Cardiovascular Control. New York, Oxford: Oxford University Press.Google Scholar
Ruggiero, D. A., Cravo, S. L., Arango, V. and Reis, D. J. (1989). Central control of the circulation by the rostral ventrolateral reticular nucleus: anatomical substrates. Prog. Brain Res., 81, 49–79.CrossRefGoogle ScholarPubMed
Rushmer, R. F. and Smith, O. A. Jr. (1959). Cardiac control. Physiol. Rev., 39, 41–68.CrossRefGoogle ScholarPubMed
Rybak, I. A., Paton, J. F. R., Rogers, R. F. and St.-John, W. M. (2002). Generation of the respiratory rhythm: state-dependency and switching. Neurocomputing, 44–46, 605–614.CrossRefGoogle Scholar
Rybak, I. A., Shevtsova, N. A., St.-John, W. M., Paton, J. F. and Pierrefiche, O. (2003). Endogenous rhythm generation in the pre-Bötzinger complex and ionic currents: modelling and in vitro studies. Eur. J. Neurosci., 18, 239–257.CrossRefGoogle ScholarPubMed
Rybak, I. A., Shevtsova, N. A., Paton, J. F.et al. (2004). Modeling the ponto-medullary respiratory network. Respir. Physiol. Neurobiol., 143, 307–319.CrossRefGoogle ScholarPubMed
Sanders, K. M. and Smith, T. K. (2003). Neural regulation of colonic motor function. In Textbook of Colonic Disease, ed. Koch, T.. Totowa, New Jersey, USA: Humana Press, Inc., pp. 35–52.CrossRefGoogle Scholar
Sanders, K. M., Ordog, T., Koh, S. D., Torihashi, S. and Ward, S. M. (1999). Development and plasticity of interstitial cells of Cajal. Neurogastroenterol. Motil., 11, 311–338.CrossRefGoogle ScholarPubMed
Sanders, K. M., Ordog, T., Koh, S. D. and Ward, S. M. (2000). A novel pacemaker mechanism drives gastrointestinal rhythmicity. News Physiol. Sci., 15, 291–298.Google ScholarPubMed
Santicioli, P. and Maggi, C. A. (1998). Myogenic and neurogenic factors in the control of pyeloureteral motility and ureteral peristalsis. Pharmacol. Rev., 50, 683–722.Google ScholarPubMed
Saper, C. B. (1995). Central autonomic system. In The Rat Nervous System, ed. Paxinos, G.. San Diego: Academic Press, pp. 107–135.Google Scholar
Saper, C. B. (2002). The central autonomic nervous system: conscious visceral perception and autonomic pattern generation. Annu. Rev. Neurosci., 25, 433–469.CrossRefGoogle ScholarPubMed
Saper, C. B. (2004). Anatomy of hypothalamus. In The Human Nervous System, ed. Paxinos, G. and Mai, J. K.. Amsterdam: Elsevier Academic Press, pp. 513–550.Google Scholar
Saphier, D. (1993). Psychoimmunology: the missing link. In Hormonally Induced Changes in Mind and Brain, ed. Schulkin, J.. Boston, New York: Academic Press, pp. 191–224.Google Scholar
Sato, A. (1972). Somato-sympathetic reflex discharges evoked through supramedullary pathways. Pflügers Arch., 332, 117–126.CrossRefGoogle ScholarPubMed
Sato, A. and Schmidt, R. F. (1971). Spinal and supraspinal components of the reflex discharges into lumbar and thoracic white rami. J. Physiol., 212, 839–850.CrossRefGoogle ScholarPubMed
Sato, A. and Schmidt, R. F. (1973). Somatosympathetic reflexes: afferent fibers, central pathways, discharge characteristics. Physiol. Rev., 53, 916–947.CrossRefGoogle ScholarPubMed
Sato, A., Sato, Y. and Schmidt, R. F. (1997). The impact of somatosensory input on autonomic functions. Rev. Physiol. Biochem. Pharmacol., 130, 1–328.CrossRefGoogle ScholarPubMed
Schaible, H. G. and Schmidt, R. F. (1988). Time course of mechanosensitivity changes in articular afferents during a developing experimental arthritis. J. Neurophysiol., 60, 2180–2195.CrossRefGoogle ScholarPubMed
Schelegle, E. S. and Green, J. F. (2001). An overview of the anatomy and physiology of slowly adapting pulmonary stretch receptors. Respir. Physiol., 125, 17–31.CrossRefGoogle ScholarPubMed
Schläfke, M. E. and Loeschke, H. H. (1967). Lokalisation eines an der Regulation von Atmung und Kreislauf beteiligten Gebietes an der ventralen Oberfläche der Medulla oblongata durch Kälteblockade [Localization of a center on the ventral surface of the medulla oblongata involved in regulation of respiratory and cardiovascular system by cold blockade]. Pflügers Arch., 297, 201–220.CrossRefGoogle Scholar
Schmelz, M., Michael, K., Weidner, C.et al. (2000). Which nerve fibers mediate the axon reflex flare in human skin?NeuroReport, 11, 645–648.CrossRefGoogle ScholarPubMed
Schmidt, R., Schmelz, M., Forster, C.et al. (1995). Novel classes of responsive and unresponsive C nociceptors in human skin. J. Neurosci., 15, 333–341.CrossRefGoogle ScholarPubMed
Schmidt, R., Schmelz, M., Torebjörk, H. E. and Handwerker, H. O. (2000). Mechano-insensitive nociceptors encode pain evoked by tonic pressure to human skin. Neuroscience, 98, 793–800.CrossRefGoogle ScholarPubMed
Schmidt, R., Schmelz, M., Weidner, C., Handwerker, H. O. and Torebjörk, H. E. (2002). Innervation territories of mechano-insensitive C nociceptors in human skin. J. Neurophysiol., 88, 1859–1866.CrossRefGoogle ScholarPubMed
Schmidt-Vanderheyden, W. and Koepchen, H. P. (1967). Zum Mechanismus der Adrenalindilatation der Skeletmuskelgefäße [On the mechanism of adrenaline dilation of skeletal muscle blood vessels]. Pflügers Arch., 298, 1–11.CrossRefGoogle Scholar
Schomburg, E. D. (1990). Spinal sensorimotor systems and their supraspinal control. Neurosci. Res., 7, 265–340.CrossRefGoogle ScholarPubMed
Schramm, L. P., Strack, A. M., Platt, K. B. and Loewy, A. D. (1993). Peripheral and central pathways regulating the kidney: a study using pseudorabies virus. Brain Res., 616, 251–262.CrossRefGoogle ScholarPubMed
Schreihofer, A. M. and Guyenet, P. G. (1997). Identification of C1 presympathetic neurons in rat rostral ventrolateral medulla by juxtacellular labeling in vivo. J. Comp. Neurol., 387, 524–536.3.0.CO;2-4>CrossRefGoogle ScholarPubMed
Schreihofer, A. M. and Guyenet, P. G. (2000). Sympathetic reflexes after depletion of bulbospinal catecholaminergic neurons with anti-dopamine-β-hydroxylase-saporin. Am. J. Physiol. Regul. Integr. Comp. Physiol., 279, R729–R742.CrossRefGoogle Scholar
Schreihofer, A. M. and Guyenet, P. G. (2002). The baroreflex and beyond: control of sympathetic vasomotor tone by γ-aminobutyric acidergic neurons in the ventrolateral medulla. Clin. Exp. Pharmacol. Physiol., 29, 514–521.CrossRefGoogle Scholar
Schreihofer, A. M. and Guyenet, P. G. (2003). Baro-activated neurons with pulse-modulated activity in the rat caudal ventrolateral medulla express GAD67 messenger ribonucleic acid. J. Neurophysiol., 89, 1265–1277.CrossRefGoogle Scholar
Schreihofer, A. M., Stornetta, R. L. and Guyenet, P. G. (1999). Evidence for glycinergic respiratory neurons: Bötzinger neurons express messenger ribonucleic acid for glycinergic transporter 2. J. Comp. Neurol., 407, 583–597.3.0.CO;2-E>CrossRefGoogle ScholarPubMed
Schreihofer, A. M., Stornetta, R. L. and Guyenet, P. G. (2000). Regulation of sympathetic tone and arterial pressure by rostral ventrolateral medulla after depletion of C1 cells in rat. J. Physiol., 529, 221–236.CrossRefGoogle ScholarPubMed
Schulkin, J. (2003a). Allostasis: a neural behavioral perspective. Horm. Behav., 43, 21–27.CrossRefGoogle Scholar
Schulkin, J. (2003b). Rethinking Homeostasis. Allostatic Regulation in Physiology and Pathophysiology. Cambridge Massachusetts: The MIT Press.Google Scholar
Schwartz, M. W. and Morton, G. J. (2002). Obesity: keeping hunger at bay. Nature, 418, 595–597.CrossRefGoogle ScholarPubMed
Schwarzacher, S. W., Wilhelm, Z., Anders, K. and Richter, D. W. (1991). The medullary respiratory network in the rat. J. Physiol., 435, 631–644.CrossRefGoogle ScholarPubMed
Seagard, J. L., Gallenberg, L. A., Hopp, F. A. and Dean, C. (1992). Acute resetting in two functionally different types of carotid baroreceptors. Circ. Res., 70, 559–565.CrossRefGoogle ScholarPubMed
Seals, D. R., Suwarno, N. O. and Dempsey, J. A. (1990). Influence of lung volume on sympathetic nerve discharge in normal humans. Circ. Res., 67, 130–141.CrossRefGoogle ScholarPubMed
Seals, D. R., Suwarno, N. O., Joyner, M. J.et al. (1993). Respiratory modulation of muscle sympathetic nerve activity in intact and lung denervated humans. Circ. Res., 72, 440–454.CrossRefGoogle ScholarPubMed
Seller, H., Langhorst, P., Richter, D. and Koepchen, H. P. (1968). Über die Abhängigkeit der pressoreceptorischen Hemmung des Sympathicus von der Atemphase und ihre Auswirkung in der Vasomotorik [Respiratory variations of baroreceptor reflex transmission and their effects on sympathetic activity and vasomotor tone]. Pflügers Arch., 302, 300–314.CrossRefGoogle Scholar
Selyanko, A. A. (1992). Membrane properties and firing characteristics of rat cardiac neurones in vitro. J. Auton. Nerv. Syst., 39, 181–189.CrossRefGoogle ScholarPubMed
Semans, J. H. and Langworthy, O. R. (1938). Observations on the neurophysiology of sexual function in the male cat. J. Urol., 40, 836–846.CrossRefGoogle Scholar
Sengupta, J. N. and Gebhart, G. F. (1994a). Characterization of mechanosensitive pelvic nerve afferent fibers innervating the colon of the rat. J. Neurophysiol., 71, 2046–2060.CrossRefGoogle Scholar
Sengupta, J. N. and Gebhart, G. F. (1994b). Mechanosensitive properties of pelvic nerve afferent fibers innervating the urinary bladder of the rat. J. Neurophysiol., 72, 2420–2430.CrossRefGoogle Scholar
Sengupta, J. N. and Gebhart, G. F. (1995). Mechanosensitive afferent fibers in the gastrointestinal and lower urinary tracts. In Visceral Pain, ed. Gebhart, G. F.. Seattle: IASP Press, pp. 75–98.Google Scholar
Sengupta, J. N., Saha, J. K. and Goyal, R. K. (1990). Stimulus–response function studies of esophageal mechanosensitive nociceptors in sympathetic afferents of opossum. J. Neurophysiol., 64, 796–812.CrossRefGoogle ScholarPubMed
Sengupta, J. N., Su, X. and Gebhart, G. F. (1996). Kappa, but not mu or delta, opioids attenuate responses to distention of afferent fibers innervating the rat colon. Gastroenterology, 111, 968–980.CrossRefGoogle ScholarPubMed
Shade, R. E., Bishop, V. S., Haywood, J. R. and Hamm, C. K. (1990). Cardiovascular and neuroendocrine responses to baroreceptor denervation in baboons. Am. J. Physiol., 258, R930–R938.Google ScholarPubMed
Shafton, A. D., Oldfield, B. J. and McAllen, R. M. (1992). CRF-like immunoreactivity selectively labels preganglionic sudomotor neurons in cat. Brain Res., 599, 253–260.CrossRefGoogle ScholarPubMed
Shah, S. D., Tse, T. F., Clutter, W. E. and Cryer, P. E. (1984). The human sympathochromaffin system. Am. J. Physiol., 247, E380–E384.Google ScholarPubMed
Shanahan, F. (1994). The intestinal immune system. In Physiology of the Gastrointestinal Tract, 3rd edn., ed. Johnson, L. R.. New York: Raven Press, pp. 643–684.Google Scholar
Sharkey, K. A. and Mawe, G. M. (2002). Neuroimmune and epithelial interactions in intestinal inflammation. Curr. Opin. Pharmacol., 2, 669–677.CrossRefGoogle ScholarPubMed
Sheehan, D. (1936). Discovery of the autonomic nervous system. Arch. Neurol. Psychiat., 35, 1081–1115.CrossRefGoogle Scholar
Sheehan, D. (1941). The autonomic nervous system prior to Gaskell. New Engl. J. Med., 224, 457–460.CrossRefGoogle Scholar
Shefchyk, S. J. (2001). Sacral spinal interneurons and the control of urinary bladder and urethral striated sphincter muscle function. J. Physiol., 533, 57–63.CrossRefGoogle ScholarPubMed
Shefchyk, S. J. (2002). Spinal cord neural organization controlling the urinary bladder and striated sphincter. Prog. Brain Res., 137, 71–82.CrossRefGoogle ScholarPubMed
Shepherd, J. Z. and Vatner, S. F. (eds.) (1996). Nervous Control of the Heart. Vol. 9, The Autonomic Nervous System, Amsterdam: Harwood Academic Publishers.Google Scholar
Sherbourne, C. D., Gonzales, R., Goldyne, M. E. and Levine, J. D. (1992). Norepinephrine-induced increase in sympathetic neuron-derived prostaglandins is independent of neuronal release mechanisms. Neurosci. Lett., 139, 188–190.CrossRefGoogle ScholarPubMed
Sherrington, C. S. (1900). Cutaneous sensation. In Textbook of Physiology, Vol. 2, ed. Schäfer, E. A.. Edinburgh, London: Young J. Pentland, pp. 920–1001.Google Scholar
Sherrington, C. (1947) [1906]. The Integrative Action of the Nervous System, 2nd edn. New Haven: Yale University Press.Google Scholar
Sillito, A. M. and Zbrozyna, A. W. (1970a). The localization of pupilloconstrictor function within the mid-brain of the cat. J. Physiol., 211, 461–477.CrossRefGoogle Scholar
Sillito, A. M. and Zbrozyna, A. W. (1970b). The activity characteristics of the preganglionic pupilloconstrictor neurones. J. Physiol., 211, 767–779.CrossRefGoogle Scholar
Silva-Carvalho, L., Paton, J. F., Goldsmith, G. E. and Spyer, K. M. (1991). The effects of electrical stimulation of lobule IXb of the posterior cerebellar vermis on neurones within the rostral ventrolateral medulla in the anaesthetised cat. J. Auton. Nerv. Syst., 36, 97–106.CrossRefGoogle ScholarPubMed
Silverberg, A. B., Shah, S. D., Haymond, M. W. and Cryer, P. E. (1978). Norepinephrine: hormone and neurotransmitter. Am. J. Physiol., 234, E252–E256.Google ScholarPubMed
Simon, E. (1974). Temperature regulation: the spinal cord as a site of extrahypothalamic thermoregulatory functions. Rev. Physiol. Biochem. Pharmacol., 71, 1–76.CrossRefGoogle Scholar
Simon, E., Pierau, F. K. and Taylor, D. C. (1986). Central and peripheral thermal control of effectors in homeothermic temperature regulation. Physiol. Rev., 66, 235–300.CrossRefGoogle ScholarPubMed
Simone, D. A., Sorkin, L. S., Oh, U.et al. (1991). Neurogenic hyperalgesia: central neural correlates in responses of spinothalamic tract neurons. J. Neurophysiol., 66, 228–246.CrossRefGoogle ScholarPubMed
Sjövall, H., Jodal, M. and Lundgren, O. (1987). Sympathetic control of intestinal fluid and electrolyte transport. News Physiol. Sci., 2, 214–217.Google Scholar
Skok, V. I. (1986). Spontaneous and reflex activities: general characteristics. In Autonomic and Enteric Ganglia, ed. Karczmar, K., Koketsu, K. and Nishi, S.. New York, London: Plenum Press, pp. 425–438.CrossRefGoogle Scholar
Skok, V. I. (2002). Nicotinic acetylcholine receptors in autonomic ganglia. Auton. Neurosci., 97, 1–11.CrossRefGoogle ScholarPubMed
Skok, V. I. and Ivanov, A. Y. (1983). What is the ongoing activity of sympathetic neurons?J. Auton. Nerv. Syst., 7, 263–270.CrossRefGoogle ScholarPubMed
Skok, V. I. and Ivanov, A. Y. (1987). Organization of presynaptic input to neurones of a sympathetic ganglion. In Organization of the Autonomic Nervous System: Central and Peripheral Mechanisms, ed. Ciriello, J., Calaresu, F. R., Renaud, L. P. and Polosa, C.. New York: Alan Liss, pp. 37–46.Google Scholar
Skok, V. I. and Ivanov, A. Y. (1989). ectectbehhaя aktиbhoctь beΓetatиbhьix ΓahΓлиeb [Natural activity in autonomic ganglia]. Kiev: Publisher Naukova Duma.
Sly, D. J., Colvill, L., McKinley, M. J. and Oldfield, B. J. (1999). Identification of neural projections from the forebrain to the kidney, using the virus pseudorabies. J. Auton. Nerv. Syst., 77, 73–82.CrossRefGoogle ScholarPubMed
Smeyne, R. J., Klein, R., Schnapp, A.et al. (1994). Severe sensory and sympathetic neuropathies in mice carrying a disrupted Tr/nerve growth factor receptor gene. Nature, 368, 246–249.CrossRefGoogle Scholar
Smith, J. C., Ellenberger, H. H., Ballanyi, K., Richter, D. W. and Feldman, J. L. (1991). Pre-Bötzinger complex: a brain stem region that may generate respiratory rhythm in mammals. Science, 254, 726–729.CrossRefGoogle ScholarPubMed
Smith, J. C., Butera, R. J., Koshiya, N.et al. (2000). Respiratory rhythm generation in neonatal and adult mammals: the hybrid pacemaker-network model. Respir. Physiol., 122, 131–147.CrossRefGoogle ScholarPubMed
Smith, J. E., Jansen, A. S., Gilbey, M. P. and Loewy, A. D. (1998). central nervous system cell groups projecting to sympathetic outflow of tail artery: neural circuits involved in heat loss in the rat. Brain Res., 786, 153–164.CrossRefGoogle ScholarPubMed
Smith, O. A. and DeVito, J. L. (1984). Central neural integration for the control of autonomic responses associated with emotion. Annu. Rev. Neurosci., 7, 43–65.CrossRefGoogle ScholarPubMed
Smith, O. A., Hohimer, A. R., Astley, C. A. and Taylor, D. J. (1979). Renal and hindlimb vascular control during acute emotion in the baboon. Am. J. Physiol., 236, R198–R205.Google ScholarPubMed
Smith, O. A., Astley, C. A., DeVito, J. L., Stein, J. M. and Walsh, K. E. (1980). Functional analysis of hypothalamic control of the cardiovascular responses accompanying emotional behavior. Fed. Proc., 39, 2487–2494.Google ScholarPubMed
Smith, O. A., Astley, C. A., Spelman, F. A.et al. (2000). Cardiovascular responses in anticipation of changes in posture and locomotion. Brain Res. Bull., 53, 69–76.CrossRefGoogle ScholarPubMed
Smith, R. E. and Horwitz, B. A. (1969). Brown fat and thermogenesis. Physiol. Rev., 49, 330–425.CrossRefGoogle ScholarPubMed
Smith, T. K., Oliver, G. R., Hennig, G. W.et al. (2003). A smooth muscle tone-dependent stretch-activated migrating motor pattern in isolated guinea-pig distal colon. J. Physiol., 551, 955–969.CrossRefGoogle ScholarPubMed
Söderholm, J. D. and Perdue, M. H. (2001). Stress and gastrointestinal tract. II. Stress and intestinal barrier function. Am. J. Physiol. Gastrointest. Liver Physiol., 280, G7–G13.CrossRefGoogle ScholarPubMed
Sonnenschein, R. R. and Weissman, M. L. (1978). Sympathetic vasomotor outflows to hindlimb muscles of the cat. Am. J. Physiol., 235, H482–H487.Google ScholarPubMed
Spencer, N. J. and Smith, T. K. (2001). Simultaneous intracellular recordings from longitudinal and circular muscle during the peristaltic reflex in guinea-pig distal colon. J. Physiol., 533, 787–799.CrossRefGoogle ScholarPubMed
Spencer, N. J. and Smith, T. K. (2004). Mechanosensory S-neurons rather than AH-neurons appear to generate a rhythmic motor pattern in guinea-pig distal colon. J. Physiol., 558, 577–596.CrossRefGoogle ScholarPubMed
Spencer, N., Walsh, M. and Smith, T. K. (1999). Does the guinea-pig ileum obey the ‘law of the intestine’? J. Physiol., 517, 889–898.CrossRefGoogle ScholarPubMed
Spencer, N. J., Hennig, G. W. and Smith, T. K. (2002). A rhythmic motor pattern activated by circumferential stretch in guinea-pig distal colon. J. Physiol., 545, 629–648.CrossRefGoogle ScholarPubMed
Spencer, N. J., Hennig, G. W. and Smith, T. K. (2003a). Stretch-activated neuronal pathways to longitudinal and circular muscle in guinea pig distal colon. Am. J. Physiol. Gastrointest. Liver Physiol., 284, G231–G241.CrossRefGoogle Scholar
Spencer, N. J., Sanders, K. M. and Smith, T. K. (2003b). Migrating motor complexes do not require electrical slow waves in the mouse small intestine. J. Physiol., 553, 881–893.CrossRefGoogle Scholar
Spencer, S. E., Sawyer, W. B., Wada, H., Platt, K. B. and Loewy, A. D. (1990). central nervous system projections to the pterygopalatine parasympathetic preganglionic neurons in the rat: a retrograde transneuronal viral cell body labeling study. Brain Res., 534, 149–169.CrossRefGoogle ScholarPubMed
Spike, R. C., Puskar, Z., Andrew, D. and Todd, A. J. (2003). A quantitative and morphological study of projection neurons in lamina I of the rat lumbar spinal cord. Eur. J. Neurosci., 18, 2433–2448.CrossRefGoogle ScholarPubMed
Spillane, D. M. (1981). The Doctrine of the Nerves. Oxford: Oxford University Press.Google Scholar
Spyer, K. M. (1981). Neural organisation and control of the baroreceptor reflex. Rev. Physiol. Biochem. Pharmacol., 88, 24–124.Google ScholarPubMed
Spyer, K. M. (1994). Central nervous mechanisms contributing to cardiovascular control. J. Physiol., 474, 1–19.CrossRefGoogle ScholarPubMed
Squire, L. R., Bloom, F. E., McConnell, S. R.et al. (eds.) (2003). Fundamental Neuroscience, 2nd edn. San Diego: Academic Press.Google Scholar
St.-John, W. M. (1998). Neurogenesis of patterns of automatic ventilatory activity. Prog. Neurobiol., 56, 97–117.CrossRefGoogle ScholarPubMed
St.-John, W. M. and Paton, J. F. (2003). Defining eupnea. Respir. Physiol. Neurobiol., 139, 97–103.CrossRefGoogle ScholarPubMed
St.-John, W. M. and Paton, J. F. (2004). Role of pontile mechanisms in the neurogenesis of eupnea. Respir. Physiol. Neurobiol., 143, 321–332.CrossRefGoogle ScholarPubMed
Stanton-Hicks, M., Jänig, W., Hassenbusch, S.et al. (1995). Reflex sympathetic dystrophy: changing concepts and taxonomy. Pain, 63, 127–133.CrossRefGoogle ScholarPubMed
Staras, K., Chang, H. S. and Gilbey, M. P. (2001). Resetting of sympathetic rhythm by somatic afferents causes post-reflex coordination of sympathetic activity in rat. J. Physiol., 533, 537–545.CrossRefGoogle ScholarPubMed
Stein, R. D. and Weaver, L. C. (1988). Multi- and single-fibre mesenteric and renal sympathetic responses to chemical stimulation of intestinal receptors in cats. J. Physiol., 396, 155–172.CrossRefGoogle ScholarPubMed
Sterling, P. and Eyer, J. (1988). Allostasis: a new paradigm to explain arousal pathology. In Handbook of Life Stress, Cognition and Health, ed. Fisher, S. and Reason, J.. New York: Wiley, pp. 629–649.Google Scholar
Stjernberg, L. and Wallin, B. G. (1983). Sympathetic neural outflow in spinal man. A preliminary report. J. Auton. Nerv. Syst., 7, 313–318.CrossRefGoogle Scholar
Stjernberg, L., Blumberg, H. and Wallin, B. G. (1986). Sympathetic activity in man after spinal cord injury. Outflow to muscle below the lesion. Brain, 109, 695–715.CrossRefGoogle ScholarPubMed
Stornetta, R. L., Schreihofer, A. M., Pelaez, N. M., Sevigny, C. P. and Guyenet, P. G. (2001). Preproenkephalin messenger ribonucleic acid is expressed by C1 and non-C1 barosensitive bulbospinal neurons in the rostral ventrolateral medulla of the rat. J. Comp. Neurol., 435, 111–126.CrossRefGoogle ScholarPubMed
Stornetta, R. L., Sevigny, C. P., Schreihofer, A. M., Rosin, D. L. and Guyenet, P. G. (2002). Vesicular glutamate transporter DNPI/VGLUT2 is expressed by both C1 adrenergic and nonaminergic presympathetic vasomotor neurons of the rat medulla. J. Comp. Neurol., 444, 207–220.CrossRefGoogle ScholarPubMed
Stornetta, R. L., McQuiston, T. J. and Guyenet, P. G. (2004). γ-aminobutyric acidergic and glycinergic presympathetic neurons of rat medulla oblongata identified by retrograde transport of pseudorabies virus and in situ hybridization. J. Comp. Neurol., 479, 257–270.CrossRefGoogle Scholar
Strack, A. M. and Loewy, A. D. (1990). Pseudorabies virus: a highly specific transneuronal cell body marker in the sympathetic nervous system. J. Neurosci., 10, 2139–2147.CrossRefGoogle ScholarPubMed
Strack, A. M., Sawyer, W. B., Marubio, L. M. and Loewy, A. D. (1988). Spinal origin of sympathetic preganglionic neurons in the rat. Brain Res., 455, 187–191.CrossRefGoogle ScholarPubMed
Strack, A. M., Sawyer, W. B., Hughes, J. H., Platt, K. B. and Loewy, A. D. (1989a). A general pattern of central nervous system innervation of the sympathetic outflow demonstrated by transneuronal pseudorabies viral infections. Brain Res., 491, 156–162.CrossRefGoogle Scholar
Strack, A. M., Sawyer, W. B., Platt, K. B. and Loewy, A. D. (1989b). central nervous system cell groups regulating the sympathetic outflow to adrenal gland as revealed by transneuronal cell body labeling with pseudorabies virus. Brain Res., 491, 274–296.CrossRefGoogle Scholar
Strauther, G. R., Longo, W. E., Virgo, K. S. and Johnson, F. E. (1999). Appendicitis in patients with previous spinal cord injury. Am. J. Surg., 178, 403–405.CrossRefGoogle ScholarPubMed
Stricker, E. M. and Verbalis, J. G. (2003). Water intake and body fluids. In Fundamental Neuroscience, 2nd edn., ed. Squire, L. R., Bloom, F. E., McConnell, S. K.et al. San Diego: Academic Press, pp. 1011–1029.Google Scholar
Strickland, J. H. and Calhoun, M. L. (1963). The integumentary system of the cat. Am. J. Vet. Res., 24, 1019–1029.Google ScholarPubMed
Su, X. and Gebhart, G. F. (1998). Mechanosensitive pelvic nerve afferent fibers innervating the colon of the rat are polymodal in character. J. Neurophysiol., 80, 2632–2644.CrossRefGoogle ScholarPubMed
Su, X., Sengupta, J. N. and Gebhart, G. F. (1997a). Effects of kappa opioid receptor-selective agonists on responses of pelvic nerve afferents to noxious colorectal distension. J. Neurophysiol., 78, 1003–1012.CrossRefGoogle Scholar
Su, X., Sengupta, J. N. and Gebhart, G. F. (1997b). Effects of opioids on mechanosensitive pelvic nerve afferent fibers innervating the urinary bladder of the bat. J. Neurophysiol., 77, 1566–1580.CrossRefGoogle Scholar
Sugenoya, J., Iwase, S., Mano, T.et al. (1998). Vasodilator component in sympathetic nerve activity destined for the skin of the dorsal foot of mildly heated humans. J. Physiol., 507, 603–610.CrossRefGoogle ScholarPubMed
Sugiura, Y., Terui, N. and Hosoa, Y. (1989). Differences in the distribution of central terminals between visceral and somatic unmyelinated primary afferent fibers. J. Neurophysiol., 62, 834–847.CrossRefGoogle Scholar
Sun, M. K. (1995). Central neural organization and control of sympathetic nervous system in mammals. Prog. Neurobiol., 47, 157–233.CrossRefGoogle ScholarPubMed
Sun, M. K. and Guyenet, P. G. (1985). γ-aminobutyric acid-mediated baroreceptor inhibition of reticulospinal neurons. Am. J. Physiol., 249, R672–R680.Google Scholar
Sun, M. K. and Reis, D. J. (1996). Medullary vasomotor activity and hypoxic sympathoexcitation in pentobarbital-anesthetized rats. Am. J. Physiol., 270, R348–R355.Google ScholarPubMed
Sun, M. K. and Spyer, K. M. (1991). Nociceptive inputs into rostral ventrolateral medulla-spinal vasomotor neurones in rats. J. Physiol., 436, 685–700.CrossRefGoogle ScholarPubMed
Sun, M. K., Hackett, J. T. and Guyenet, P. G. (1988a). Sympathoexcitatory neurons of rostral ventrolateral medulla exhibit pacemaker properties in the presence of a glutamate-receptor antagonist. Brain Res., 438, 23–40.CrossRefGoogle Scholar
Sun, M. K., Young, B. S., Hackett, J. T. and Guyenet, P. G. (1988b). Reticulospinal pacemaker neurons of the rat rostral ventrolateral medulla with putative sympathoexcitatory function: an intracellular study in vitro. Brain Res., 442, 229–239.CrossRefGoogle Scholar
Sun, W. and Panneton, W. M. (2002). The caudal pressor area of the rat: its precise location and projections to the ventrolateral medulla. Am. J. Physiol. Regul. Integr. Comp. Physiol., 283, R768–R778.CrossRefGoogle ScholarPubMed
Sun, W. and Panneton, W. M. (2004). Defining projections from the caudal pressor area of the caudal ventrolateral medulla. J. Comp. Neurol., 482, 273–293.CrossRefGoogle Scholar
Sundlöf, G. and Wallin, B. G. (1977). The variability of muscle nerve sympathetic activity in resting recumbent man. J. Physiol., 272, 383–397.CrossRefGoogle ScholarPubMed
Sundlöf, G. and Wallin, B. G. (1978). Effect of lower body negative pressure on human muscle nerve sympathetic activity. J. Physiol., 278, 525–532.CrossRefGoogle ScholarPubMed
Sved, A. F. (2004). Tonic glutamatergic drive of rostral ventrolateral medulla vasomotor neurons? Am. J. Physiol. Regul. Integr. Comp. Physiol., 287, R1301–R1303.CrossRefGoogle Scholar
Sved, A. F., Mancini, D. L., Graham, J. C., Schreihofer, A. M. and Hoffman, G. E. (1994). PNMT-containing neurons of the C1 cell group express c-fos in response to changes in baroreceptor input. Am. J. Physiol., 266, R361–R367.Google ScholarPubMed
Swanson, L. W. (1987). The hypothalamus. In Handbook of Chemical Anatomy, Vol. 5, Integrated Systems of the central nervous system, part I: Hypothalamus, Hippocampus, Amygdala, Retina., ed. Björklund, A., Hökfelt, T. and Swanson, L. W., Amsterdam, New York, Oxford: Elsevier, pp. 1–124.Google Scholar
Swanson, L. W. (1995). Mapping the human brain: past, present, and future. Trends Neurosci., 18, 471–474.CrossRefGoogle ScholarPubMed
Swanson, L. W. (2000). Cerebral hemisphere regulation of motivated behavior. Brain Res., 886, 113–164.CrossRefGoogle ScholarPubMed
Swanson, L. W. (2003). The architecture of the nervous system. In Fundamental Neuroscience, 2nd edn., ed. Squire, L. R., Bloom, F. E., McConnell, S. K.et al. San Diego: Academic Press, pp. 15–45.Google Scholar
Swift, J. Q., Garry, M. G., Roszkowski, M. T. and Hargreaves, K. M. (1993). Effect of flurbiprofen on tissue levels on immunoreactive bradykinin and acute postoperative pain. J. Oral. Maxillofac. Surg., 51, 112–117.CrossRefGoogle ScholarPubMed
Szurszewski, J. H. (1981). Physiology of mammalian prevertebral ganglia. Annu. Rev. Physiol., 43, 53–68.CrossRefGoogle ScholarPubMed
Szurszewski, J. H. and King, B. F. (1989). Physiology of prevertebral ganglia in mammals with special reference to the inferior mesenteric ganglion. In The Gastrointestinal System, ed. Wood, J. D.. Bethesda: American Physiological Society, pp. 519–592.Google Scholar
Taché, Y., Martinez, V., Million, M. and Wang, L. (2001). Stress and the gastrointestinal tract III. Stress-related alterations of gut motor function: role of brain corticotropin-releasing factor receptors. Am. J. Physiol. Gastrointest. Liver Physiol., 280, G173–G177.CrossRefGoogle ScholarPubMed
Taiwo, Y. O. and Levine, J. D. (1988). Characterization of the arachidonic acid metabolites mediating bradykinin and noradrenaline hyperalgesia. Brain Res., 458, 402–406.CrossRefGoogle ScholarPubMed
Tanaka, M., Nagashima, K., McAllen, R. M. and Kanosue, K. (2002). Role of the medullary raphe in thermoregulatory vasomotor control in rats. J. Physiol., 540, 657–664.CrossRefGoogle ScholarPubMed
Tang, X., Neckel, N. D. and Schramm, L. P. (2003). Locations and morphologies of sympathetically correlated neurons in the T(10) spinal segment of the rat. Brain Res., 976, 185–193.CrossRefGoogle Scholar
Tang, X., Neckel, N. D. and Schramm, L. P. (2004). Spinal interneurons infected by renal injection of pseudorabies virus in the rat. Brain Res., 1004, 1–7.CrossRefGoogle ScholarPubMed
Tatarchenko, L. A., Ivanov, A. Y. and Skok, V. I. (1990). Organization of the tonically active pathways through the superior cervical ganglion of the rabbit. J. Auton. Nerv. Syst., 30 Suppl., S163–S168.CrossRefGoogle ScholarPubMed
Taylor, R. B. and Weaver, L. C. (1993). Dorsal root afferent influences on tonic firing of renal and mesenteric sympathetic nerves in rats. Am. J. Physiol., 264, R1193–R1199.Google ScholarPubMed
Horst, G. J., Hautvast, R. W., Jongste, M. J. and Korf, J. (1996). Neuroanatomy of cardiac activity-regulating circuitry: a transneuronal retrograde viral labelling study in the rat. Eur. J. Neurosci., 8, 2029–2041.CrossRefGoogle ScholarPubMed
Thompson, R. H. and Swanson, L. W. (2003). Structural characterization of a hypothalamic visceromotor pattern generator network. Brain Res. Brain Res. Rev., 41, 153–202.CrossRefGoogle ScholarPubMed
Thorén, P. (1979). Role of cardiac vagal C-fibers in cardiovascular control. Rev. Physiol. Biochem. Pharmacol., 86, 1–94.CrossRefGoogle ScholarPubMed
Thrasher, T. N. (2002). Unloading of arterial baroreceptors causes neurogenic hypertension. Am. J. Physiol. Regul. Integr. Comp. Physiol., 282, R1044–R1053.CrossRefGoogle ScholarPubMed
Thrasher, T. N. (2004). Baroreceptors and the long-term control of blood pressure. Exp. Physiol., 89, 331–335.CrossRefGoogle ScholarPubMed
Thrasher, T. N. (2005). Baroreceptors, baroreceptor unloading, and the long-term control of blood pressure. Am. J. Physiol. Integr. Comp. Physiol., 288, R819–R827.CrossRefGoogle ScholarPubMed
Thuneberg, L. (1982). Interstitial cells of Cajal: intestinal pacemaker cells? Adv. Anat. Embryol. Cell Biol., 71, 1–130.CrossRefGoogle ScholarPubMed
Thuneberg, L. (1999). One hundred years of interstitial cells of Cajal. Microsc. Res. Tech., 47, 223–238.3.0.CO;2-C>CrossRefGoogle ScholarPubMed
Tokimasa, T. and Akasu, T. (1995). Biochemical gating for voltage-gated channels: mechanisms for slow synaptic potentials in autonomic ganglia. In The Autonomic Nervous System, Vol. 6, Autonomic Ganglia, ed. McLachlan, E. M.. Chur Switzerland: Harwood Academic Publishers, pp. 259–295.Google Scholar
Tomita, T. (1975). Electrophysiology of mammalian smooth muscle. Prog. Biophys. Mol. Biol., 30, 185–203.CrossRefGoogle ScholarPubMed
Tomori, Z. and Widdicombe, J. G. (1969). Muscular, bronchomotor and cardiovascular reflexes elicited by mechanical stimulation of the respiratory tract. J. Physiol., 200, 25–49.CrossRefGoogle ScholarPubMed
Torebjörk, H. E., Lundberg, L. E. R. and LaMotte, R. H. (1992). Central changes in processing of mechanoreceptive input in capsaicin-induced secondary hyperalgesia in humans. J. Physiol., 448, 765–780.CrossRefGoogle ScholarPubMed
Traube, L. (1865). Über periodische Thätigkeits-Aeusserungen des vasomotorischen und Hemmungs-Nervencentrums [On the periodic changes of the vasomotor and inhibitory neural centers]. Centralbl. Medic. Wissensch., 56, 881–885.Google Scholar
Travagli, R. A. and Rogers, R. C. (2001). Receptors and transmission in the brain-gut axis: potential for novel therapies. V. Fast and slow extrinsic modulation of dorsal vagal complex circuits. Am. J. Physiol., 281, G595–G601.Google ScholarPubMed
Travagli, R. A., Hermann, G. E., Browning, K. N. and Rogers, C. (2006). Brain stem circuits regulating gastric function. Annu. Rev. Physiol., 68, 279–305.CrossRefGoogle ScholarPubMed
Treede, R. D., Kenshalo, D. R., Gracely, R. H. and Jones, A. K. (1999). The cortical representation of pain. Pain, 79, 105–111.CrossRefGoogle Scholar
Treede, R. D., Apkarian, A. V., Bromm, B., Greenspan, J. D. and Lenz, F. A. (2000). Cortical representation of pain: functional characterization of nociceptive areas near the lateral sulcus. Pain, 87, 113–119.CrossRefGoogle ScholarPubMed
Tsunoo, A., Konishi, S. and Otsuka, M. (1982). Substance P as an excitatory transmitter of primary afferent neurons in guinea-pig sympathetic ganglia. Neuroscience, 7, 2025–2037.CrossRefGoogle ScholarPubMed
Tucker, D. C., Saper, C. B., Ruggiero, D. A. and Reis, D. J. (1987). Organization of central adrenergic pathways: I. Relationships of ventrolateral medullary projections to the hypothalamus and spinal cord. J. Comp. Neurol., 259, 591–603.CrossRefGoogle ScholarPubMed
Ulman, L. G., Potter, E. K. and McCloskey, D. I. (1992). Effects of sympathetic activity and galanin on cardiac vagal action in anaesthetized cats. J. Physiol., 448, 225–235.CrossRefGoogle ScholarPubMed
Undem, B. and Weinreich, D. (eds.) (2005). Advance in Vagal Afferent Neurobiology. Boca Raton: CRC Press.CrossRefGoogle Scholar
Unzicker, K. (ed.) (1996). The Autonomic Nervous System Vol. 10, Autonomic–Endocrine Interactions. Amsterdam: Harwood Academic Publishers.Google Scholar
Uvnäs, B. (1954). Sympathetic vasodilator outflow. Pharmacol. Rev., 34, 608–618.Google ScholarPubMed
Uvnäs, B. (1960). Sympathetic vasodilator system and blood flow. Physiol. Rev., 40, Suppl. 4, 68–75.Google Scholar
Vallbo, A. B., Hagbarth, K. E., Torebjörk, H. E. and Wallin, B. G. (1979). Somatosensory, proprioceptive and sympathetic activity in human peripheral nerves. Physiol. Rev., 59, 919–957.CrossRefGoogle ScholarPubMed
Vallbo, A. B., Olsson, K. A., Westberg, K. G. and Clark, F. J. (1984). Microstimulation of single tactile afferents from the human hand. Sensory attributes related to unit type and properties of receptive fields. Brain, 107, 727–749.CrossRefGoogle ScholarPubMed
Vallbo, A. B., Hagbarth, K. E. and Wallin, B. G. (2004). Microneurography: how the technique developed and its role in the investigation of the sympathetic nervous system. J. Appl. Physiol., 96, 1262–1269.CrossRefGoogle ScholarPubMed
Bockstaele, E. J., Aston-Jones, G., Pieribone, V. A., Ennis, M. and Shipley, M. T. (1991). Subregions of the periaqueductal gray topographically innervate the rostral ventral medulla in the rat. J. Comp. Neurol., 309, 305–327.CrossRefGoogle ScholarPubMed
Dijk, J. G. (2003). Fainting in animals. Clin. Auton. Res., 13, 247–255.CrossRefGoogle ScholarPubMed
Helden, D. F. (1988a). An alpha-adrenoceptor-mediated chloride conductance in mesenteric veins of the guinea-pig. J. Physiol., 401, 489–501.CrossRefGoogle Scholar
Helden, D. F. (1988b). Electrophysiology of neuromuscular transmission in guinea-pig mesenteric veins. J. Physiol., 401, 469–488.CrossRefGoogle Scholar
Helden, D. F. (1991). Spontaneous and noradrenaline-induced transient depolarizations in the smooth muscle of guinea-pig mesenteric vein. J. Physiol., 437, 511–541.CrossRefGoogle ScholarPubMed
Vecchiet, L., Albe-Fessard, D., Lindblom, U. and Giamberardino, M. A. (eds.) (1993). New Trends in Referred Pain and Hyperalgesia. Amsterdam: Elsevier Science.Google Scholar
Verberne, A. J., Stornetta, R. L. and Guyenet, P. G. (1999). Properties of C1 and other ventrolateral medullary neurons with hypothalamic projections in the rat. J. Physiol., 517, 477–494.CrossRefGoogle ScholarPubMed
Victor, R. G., Leimbach, W. N. J., Seals, D. R., Wallin, B. G. and Mark, A. L. (1987). Effects of the cold pressor test on muscle sympathetic nerve activity in humans. Hypertension, 9, 429–436.CrossRefGoogle ScholarPubMed
Victor, R. G., Secher, N. H., Lyson, T. and Mitchell, J. H. (1995). Central command increases muscle sympathetic nerve activity during intense intermittent isometric exercise in humans. Circ. Res., 76, 127–131.CrossRefGoogle ScholarPubMed
Villanueva, L. and Nathan, P. W. (2000). Multiple pain pathways. In Proceedings of the 9th World Congress on Pain, ed. Devor, M., Rowbotham, M. C. and Weisenfeld-Hallin, Z. Seattle: IASP Press, pp. 371–386.Google Scholar
Vissing, S. F., Scherrer, U. and Victor, R. G. (1994). Increase of sympathetic discharge to skeletal muscle but not to skin during mild lower body negative pressure in humans. J. Physiol., 481, 233–241.CrossRefGoogle Scholar
Vizzard, M. A., Erickson, V. L., Card, J. P., Roppolo, J. R. and Groat, W. C. (1995). Transneuronal labeling of neurons in the adult rat brainstem and spinal cord after injection of pseudorabies virus into the urethra. J. Comp. Neurol., 355, 629–640.CrossRefGoogle ScholarPubMed
Vizzard, M. A., Brisson, M. and Groat, W. C. (2000). Transneuronal labeling of neurons in the adult rat central nervous system following inoculation of pseudorabies virus into the colon. Cell Tissue Res., 299, 9–26.CrossRefGoogle ScholarPubMed
Vogt, B. A. (2005). Pain and emotion interactions in subregions of the cingulate gyrus. Nat. Rev. Neurosci., 6, 533–544.CrossRefGoogle ScholarPubMed
Vogt, B. A., Rosene, D. L. and Pandya, D. N. (1979). Thalamic and cortical afferents differentiate anterior from posterior cingulate cortex in the monkey. Science, 204, 205–207.CrossRefGoogle ScholarPubMed
Vollmer, R. R. (1996). Selective neural regulation of epinephrine and norepinephrine cells in the adrenal medulla – cardiovascular implications. Clin. Exp. Hypertens., 18, 731–751.CrossRefGoogle ScholarPubMed
Euler, U. S. (1956). Noradrenaline: Chemistry, Physiology, Pharmacology and Clinical Aspects. Springfield Ill: Thomas.Google Scholar
von Euler, C. (1986). Brain stem mechanisms for generation and control of breathing pattern. In Handbook of Physiology, The Respiratory System, Vol. II, ed. Fishman, A. P., Cherniack, N. S. and Widdicombe, J. G.. Bethesda: American Physiological Society, pp. 1–67.Google Scholar
Holst, E. and St. Paul, U. (1960). Vom Wirkungsgefüge der Triebe [The Wirkungsgefüge of Drives]. Die Naturwissenschaften, 47, 409–422.CrossRefGoogle Scholar
Holst, E. and St. Paul, U. (1962). Electrically controlled behavior. Sci. Am., 206, 50–60.CrossRefGoogle Scholar
Voyvodic, J. T. (1989). Peripheral target regulation of dendritic geometry in the rat superior cervical ganglion. J. Neurosci., 9, 1997–2010.CrossRefGoogle ScholarPubMed
Wallin, B. G. (1990). Neural control of human skin blood flow. J. Auton. Nerv. Syst., 30 Suppl., S185–S190.CrossRefGoogle ScholarPubMed
Wallin, B. G. (2002). Intraneural recordings of normal and abnormal sympathetic activity in humans. In Autonomic Failure, 4th edn., ed. Mathias, C. J. and Bannister, R.. Oxford: Oxford University Press, pp. 224–231.Google ScholarPubMed
Wallin, B. G. and Elam, M. (1994). Insights from intraneural recordings of sympathetic nerve traffic in humans. News Physiol. Sci., 9, 203–207.Google Scholar
Wallin, B. G. and Fagius, J. (1986). The sympathetic nervous system in man – aspects derived from microelectrode recordings. Trends Neurosci., 9, 63–67.CrossRefGoogle Scholar
Wallin, B. G. and Fagius, J. (1988). Peripheral sympathetic neural activity in conscious humans. Annu. Rev. Physiol., 50, 565–576.CrossRefGoogle ScholarPubMed
Wallin, B. G. and Stjernberg, L. (1984). Sympathetic activity in man after spinal cord injury. Outflow to skin below the lesion. Brain, 107, 183–198.CrossRefGoogle ScholarPubMed
Wallin, B. G., Batelsson, K., Kienbaum, P., Karlsson, T. and Gazelius, B. (1998). Two neural mechanisms for respiration-induced cutaneous vasodilatation in humans. J. Physiol., 513, 559–569.CrossRefGoogle ScholarPubMed
Wang, F. B., Holst, M. C. and Powley, T. L. (1995). The ratio of pre- to postganglionic neurons and related issues in the autonomic nervous system. Brain Res. Brain Res. Rev., 21, 93–115.CrossRefGoogle ScholarPubMed
Wang, H., Germanson, T. P. and Guyenet, P. G. (2002). Depressor and tachypneic responses to chemical stimulation of the ventral respiratory group are reduced by ablation of neurokinin-1 receptor-expressing neurons. J. Neurosci., 22, 3755–3764.CrossRefGoogle ScholarPubMed
Wang, H., Weston, M. C., McQuiston, T. J., Stornetta, R. L. and Guyenet, P. G. (2003). Neurokinin-1 receptor-expressing cells regulate depressor region of rat ventrolateral medulla. Am. J. Physiol. Heart Circ. Physiol., 285, H2757–H2769.CrossRefGoogle ScholarPubMed
Wank, M. and Neuhuber, W. L. (2001). Local differences in vagal afferent innervation of the rat esophagus are reflected by neurochemical differences at the level of the sensory ganglia and by different brainstem projections. J. Comp. Neurol., 435, 41–59.CrossRefGoogle ScholarPubMed
Ward, S. M. and Sanders, K. M. (2001). Physiology and pathophysiology of the interstitial cell of Cajal: from bench to bedside. I. Functional development and plasticity of interstitial cells of Cajal networks. Am. J. Physiol. Gastrointest. Liver Physiol., 281, G602–G611.CrossRefGoogle ScholarPubMed
Ward, S. M., Morris, G., Reese, L., Wang, X. Y. and Sanders, K. M. (1998). Interstitial cells of Cajal mediate enteric inhibitory neurotransmission in the lower esophageal and pyloric sphincters. Gastroenterology, 115, 314–329.CrossRefGoogle ScholarPubMed
Ward, S. M., Beckett, E. A., Wang, X.et al. (2000a). Interstitial cells of Cajal mediate cholinergic neurotransmission from enteric motor neurons. J. Neurosci., 20, 1393–1403.CrossRefGoogle Scholar
Ward, S. M., Ordog, T., Koh, S. D.et al. (2000b). Pacemaking in interstitial cells of Cajal depends upon calcium handling by endoplasmic reticulum and mitochondria. J. Physiol., 525, 355–361.CrossRefGoogle Scholar
Wasner, G., Heckmann, K., Maier, C. and Baron, R. (1999). Vascular abnormalities in acute reflex sympathetic dystrophy (complex regional pain syndrome I): complete inhibition of sympathetic nerve activity with recovery. Arch. Neurol., 56, 613–620.CrossRefGoogle ScholarPubMed
Wasner, G., Schattschneider, J., Heckmann, K., Maier, C. and Baron, R. (2001). Vascular abnormalities in reflex sympathetic dystrophy (complex regional pain syndrome I): mechanisms and diagnostic value. Brain, 124, 587–599.CrossRefGoogle ScholarPubMed
Watkins, L. R. and Maier, S. F. (eds.) (1999). Cytokines and Pain. Basel, Boston, Berlin: Birkhäuser Verlag.CrossRefGoogle ScholarPubMed
Watkins, L. R. and Maier, S. F. (2000). The pain of being sick: implications of immune-to-brain communication for understanding pain. Annu. Rev. Psychol., 51, 29–57.CrossRefGoogle ScholarPubMed
Watts, A. G. and Swanson, L. W. (2002). Anatomy of motivational systems. In “Stevens” Handbook of Experimental Psychology, 3rd edn. Vol. 3, ed. Gallistel, G. R.. New York: John Wiley, pp. 563–631.Google Scholar
Weaver, L. C. and Polosa, C. (1997). Spinal cord circuits providing control of sympathetic preganglionic neurons. In The Autonomic Nervous System, Vol. 11, Central Nervous Control of Autonomic Function, ed. Jordan, D.. Amsterdam: Harwood Academic Publishers, pp. 29–61.Google Scholar
Weaver, L. C., Verghese, P., Bruce, J. C.et al. (2001). Autonomic dysreflexia and primary afferent sprouting after clip-compression injury of the rat spinal cord. J. Neurotrauma, 18, 1107–1119.CrossRefGoogle ScholarPubMed
Weaver, L. C., Marsh, D. R., Gris, D., Meakin, S. O. and Dekaban, G. A. (2002). Central mechanisms for autonomic dysreflexia after spinal cord injury. Prog. Brain Res., 137, 83–95.CrossRefGoogle ScholarPubMed
Wesselmann, U. and McLachlan, E. M. (1984). The effect of previous transection on quantitative estimates of the preganglionic neurons projecting in the cervical sympathetic trunk of the guinea-pig and the cat made by retrograde labelling of damaged axons by horseradish peroxidase. Neuroscience, 13, 1299–1309.CrossRefGoogle Scholar
Westerhaus, M. J. and Loewy, A. D. (1999). Sympathetic-related neurons in the preoptic region of the rat identified by viral transneuronal labeling. J. Comp. Neurol., 414, 361–378.3.0.CO;2-X>CrossRefGoogle ScholarPubMed
Westerhaus, M. J. and Loewy, A. D. (2001). Central representation of the sympathetic nervous system in the cerebral cortex. Brain Res., 903, 117–127.CrossRefGoogle ScholarPubMed
Weston, M., Wang, H., Stornetta, R. L., Sevigny, C. P. and Guyenet, P. G. (2003). Fos expression by glutamatergic neurons of the solitary tract nucleus after phenylephrine-induced hypertension in rats. J. Comp. Neurol., 460, 525–541.CrossRefGoogle ScholarPubMed
White, J. C. and Bland, E. F. (1948). The surgical relief of severe angina pectoris. Medicine, 27, 1–42.CrossRefGoogle ScholarPubMed
Whyment, A. D., Wilson, J. M., Renaud, L. P. and Spanswick, D. (2004). Activation and integration of bilateral γ-aminobutyric acid-mediated synaptic inputs in neonatal rat sympathetic preganglionic neurons in vitro. J. Physiol., 555, 189–203.CrossRefGoogle Scholar
Widdicombe, J. G. (1966). Action potentials in parasympathetic and sympathetic efferent fibers to the trachea and lungs of dogs and cats. J. Physiol., 186, 56–88.CrossRefGoogle Scholar
Widdicombe, J. G. (1986). Sensory innervation of the lungs and airways. Prog. Brain Res., 67, 49–64.CrossRefGoogle ScholarPubMed
Widdicombe, J. (2001). Airway receptors. Respir. Physiol., 125, 3–15.CrossRefGoogle ScholarPubMed
Widdicombe, J. (2003). Functional morphology and physiology of pulmonary rapidly adapting receptors (RARs). Anat. Rec., 270A, 2–10.CrossRefGoogle Scholar
Williams, C. L., Men, D., Clayton, E. C. and Gold, P. E. (1998). Norepinephrine release in the amygdala after systemic injection of epinephrine or escapable footshock: contribution of the nucleus of the solitary tract. Behav. Neurosci., 112, 1414–1422.CrossRefGoogle ScholarPubMed
Williams, I. R. and Kupper, T. S. (1996). Immunity at the surface: homeostatic mechanisms of the skin immune system. Life Sci., 58, 1485–1507.CrossRefGoogle ScholarPubMed
Williams, R. M., Berthoud, H. R. and Stead, R. H. (1997). Vagal afferent nerve fibers contact mast cells in rat small intestinal mucosa. NeuroImmunoModulation, 4, 266–270.CrossRefGoogle ScholarPubMed
Willis, W. D. J. (1985). The Pain System. Basel: Karger.Google ScholarPubMed
Willis, W. D. Jr. (1999). Dorsal root potentials and dorsal root reflexes: a double-edged sword. Exp. Brain Res., 124, 395–421.CrossRefGoogle ScholarPubMed
Willis, W. D. J. and Coggeshall, R. E. (2004a). Sensory Mechanisms of the Spinal Cord. Primary Afferent Neurons and the Spinal Dorsal Horn, Vol. 1, 3rd edn. New York: Kluwer Academic/Plenum Publishers.Google Scholar
Willis, W. D. J. and Coggeshall, R. E. (2004b). Sensory Mechanisms of the Spinal Cord. Ascending Sensory Tracts and Their Descending Control, Vol. 2, 3rd edn. New York: Kluwer Academic/Plenum Publishers.Google Scholar
Willis, W. D., Al Chaer, E. D., Quast, M. J. and Westlund, K. N. (1999). A visceral pain pathway in the dorsal column of the spinal cord. Proc. Natl. Acad. Sci. U. S. A., 96, 7675–7679.CrossRefGoogle ScholarPubMed
Willis, W. D. Jr., Zhang, X., Honda, C. N. and Giesler, G. J. Jr. (2001). Projections from the marginal zone and deep dorsal horn to the ventrobasal nuclei of the primate thalamus. Pain, 92, 267–276.CrossRefGoogle ScholarPubMed
Willis, W. D. Jr., Zhang, X., Honda, C. N. and Giesler, G. J. Jr. (2002). A critical review of the role of the proposed posterior part of the ventromedial nucleus (thalamus, primate) nucleus in pain. J. Pain, 3, 79–94.CrossRefGoogle Scholar
Wilson, L. B., Andrew, D. and Craig, A. D. (2002). Activation of spinobulbar lamina I neurons by static muscle contraction. J. Neurophysiol., 87, 1641–1645.CrossRefGoogle ScholarPubMed
Winkler, H. (1988). Occurrence and mechanisms of exocytosis in adrenal medulla and sympathetic nerve. In Handbook of Experimental Pharmacology, Vol. 90/I Catecholamines I, ed. Trendelenburg, U. and Weiner, N.. Berlin, Heidelberg, New York: Springer-Verlag, pp. 43–118.Google Scholar
Wood, J. D. (1994). Physiology of the enteric nervous system. In Physiology of the Gastrointestinal Tract, 3rd edn., ed. Johnson, L. R.. New York: Raven Press, pp. 423–482.Google Scholar
Wood, J. D. (2002). Enteric neuro-immunology. In The Autonomic Nervous System, Vol. 14, Innervation of the Gastrointestinal Tract, ed. Brookes, S. and Costa, M.. London, New York: Taylor and Francis, pp. 363–392.Google Scholar
Woods, S. C. and Stricker, E. M. (2003). Food intake and metabolism. In Fundamental Neuroscience, 2nd edn., eds. Squire, L. R., Bloom, F. E., McConnell, S. K.et al. San Diego: Academic Press, pp. 991–1009.Google Scholar
Woolf, C. J. (1996). Phenotypic modification of primary sensory neurons: the role of nerve growth factor in the production of persistent pain. Phil. Trans. R. Soc. London B, 351, 441–448.CrossRefGoogle ScholarPubMed
Woolf, C. J., Safieh-Garabedian, B., Ma, Q.-P., Crilly, P. and Winter, J. (1994). Nerve growth factor contributes to the generation of inflammatory sensory hypersensitivity. Neuroscience, 62, 327–331.CrossRefGoogle ScholarPubMed
Woolf, C. J., Ma, Q. P., Allchorne, A. and Poole, S. (1996). Peripheral cell types contributing to the hyperalgesic action of nerve growth factor in inflammation. J. Neurosci., 16, 2716–2723.CrossRefGoogle ScholarPubMed
Xi, X., Randall, W. C. and Wurster, R. D. (1991). Intracellular recording of spontaneous activity of canine intracardiac ganglion cells. Neurosci Lett., 128, 129–132.CrossRefGoogle ScholarPubMed
Yang, H., Kawakubo, K., Wong, H.et al. (2000a). Peripheral PYY inhibits intracisternal thyrotropin-releasing hormone-induced gastric acid secretion by acting in the brain. Am. J. Physiol. Gastrointest. Liver Physiol., 279, G575–G581.CrossRefGoogle Scholar
Yang, H., Yuan, P. Q., Wang, L. and Taché, Y. (2000b). Activation of the parapyramidal region in the ventral medulla stimulates gastric acid secretion through vagal pathways in rats. Neuroscience, 95, 773–779.CrossRefGoogle Scholar
Yang, Z. and Coote, J. H. (2003). Role of γ-aminobutyric acid and nitric oxide in the paraventricular nucleus-mediated reflex inhibition of renal sympathetic nerve activity following stimulation of right atrial receptors in the rat. Exp. Physiol., 88. 335–342.CrossRefGoogle ScholarPubMed
Yang, Z., Smith, L. and Coote, J. H. (2004). Paraventricular nucleus activation of renal sympathetic neurones is synaptically depressed by nitric oxide and glycine acting at a spinal level. Neuroscience, 124, 421–428.CrossRefGoogle Scholar
Yeoh, M., McLachlan, E. M. and Brock, J. A. (2004a). Tail arteries from chronically spinal rats have potentiated responses to nerve stimulation in vitro. J. Physiol., 556, 545–555.CrossRefGoogle Scholar
Yeoh, M., McLachlan, E. M. and Brock, J. A. (2004b). Chronic decentralization potentiates neurovascular transmission in the isolated rat tail artery, mimicking the effects of spinal transection. J. Physiol., 561, 583–596.CrossRefGoogle Scholar
Yoshimura, M., Polosa, C. and Nishi, S. (1987a). Slow excitatory postsynaptic potential and the depolarizing action of noradrenaline on sympathetic preganglionic neurons. Brain Res., 414, 138–142.CrossRefGoogle Scholar
Yoshimura, M., Polosa, C. and Nishi, S. (1987b). Slow inhibitory postsynaptic potential and the noradrenaline-induced inhibition of the cat sympathetic preganglionic neuron in vitro. Brain Res., 419, 383–386.CrossRefGoogle Scholar
Young, J. B. and Landsberg, L. (2001). Synthesis, storage, and secretion of adrenal medullary hormones: physiology and pathophysiology. In Handbook of Physiology. Section 7. The Endocrine System. Vol. IV, Coping with the Environment: Neural and Neuroendocrine Mechanisms, ed. McEwan, B. S.. Oxford, New York: Oxford University Press, pp. 3–19.Google Scholar
Yu, L. C. and Perdue, M. H. (2001). Role of mast cells in intestinal mucosal function: studies in models of hypersensitivity and stress. Immunol. Rev., 179, 61–73.CrossRefGoogle Scholar
Zagon, A. and Smith, A. D. (1993). Monosynaptic projections from the rostral ventrolateral medulla oblongata to identified sympathetic preganglionic neurons. Neuroscience, 54, 729–743.CrossRefGoogle ScholarPubMed
Zagorodnyuk, V. P. and Brookes, S. J. (2000). Transduction sites of vagal mechanoreceptors in the guinea pig esophagus. J. Neurosci., 20, 6249–6255.CrossRefGoogle ScholarPubMed
Zagorodnyuk, V. P., Chen, B. N. and Brookes, S. J. (2001). Intraganglionic laminar endings are mechano-transduction sites of vagal tension receptors in the guinea-pig stomach. J. Physiol., 534, 255–268.CrossRefGoogle ScholarPubMed
Zaretsky, D. V., Zaretskaia, M. V. and DiMicco, J. A. (2003a). Stimulation and blockade of γ-aminobutyric acid(A) receptors in the raphe pallidus: effects on body temperature, heart rate, and blood pressure in conscious rats. Am. J. Physiol. Regul. Integr. Comp. Physiol., 285, R110–R116.CrossRefGoogle Scholar
Zaretsky, D. V., Zaretskaia, M. V., Samuels, B. C., Cluxton, L. K. and DiMicco, J. A. (2003b). Microinjection of muscimol into raphe pallidus suppresses tachycardia associated with air stress in conscious rats. J. Physiol., 546, 243–250.CrossRefGoogle Scholar
Zhang, X. and Giesler, G. J. Jr. (2005). Response characterstics of spinothalamic tract neurons that project to the posterior thalamus in rats. J. Neurophysiol., 93, 2552–2564.CrossRefGoogle ScholarPubMed
Zhang, X., Fogel, R. and Renehan, W. E. (1992). Physiology and morphology of neurons in the dorsal motor nucleus of the vagus and the nucleus of the solitary tract that are sensitive to distension of the small intestine. J. Comp. Neurol., 323, 432–448.CrossRefGoogle ScholarPubMed
Zhang, X., Fogel, R. and Renehan, W. E. (1995a). Relationships between the morphology and function of gastric- and intestine-sensitive neurons in the nucleus of the solitary tract. J. Comp. Neurol., 363, 37–52.CrossRefGoogle Scholar
Zhang, X., Kostarczyk, E. and Giesler, G. J. Jr. (1995b). Spinohypothalamic tract neurons in the cervical enlargement of rats: descending axons in the ipsilateral brain. J. Neurosci., 15, 8393–8407.CrossRefGoogle Scholar
Zhang, X., Renehan, W. E. and Fogel, R. (1998). Neurons in the vagal complex of the rat respond to mechanical and chemical stimulation of the GI tract. Am. J. Physiol., 274, G331–G341.Google ScholarPubMed
Zhang, X., Honda, C. N. and Giesler, G. J. Jr. (2000a). Position of spinothalamic tract axons in upper cervical spinal cord of monkeys. J. Neurophysiol., 84, 1180–1185.CrossRefGoogle Scholar
Zhang, X., Wenk, H. N., Honda, C. N. and Giesler, G. J. Jr. (2000b). Locations of spinothalamic tract axons in cervical and thoracic spinal cord white matter in monkeys. J. Neurophysiol., 83, 2869–2880.CrossRefGoogle Scholar
Zhao, F. Y., Saito, K., Yoshioka, K.et al. (1996). Tachykininergic synaptic transmission in the coeliac ganglion of the guinea-pig. Br. J. Pharmacol., 118, 2059–2066.CrossRefGoogle ScholarPubMed
Zhu, J. X., Zhu, X. Y., Owyang, C. and Li, Y. (2001). Intestinal serotonin acts as a paracrine substance to mediate vagal signal transmission evoked by luminal factors in the rat. J. Physiol., 530, 431–442.CrossRefGoogle ScholarPubMed
Ádám, G. (1967). Interoception and Behavior: An Experimental Study. Tanslated from the Hungarian by R. de Chantel, revised by H. Slucki. Budapest: Akadémiai Kiadó.Google Scholar
Ádám, G. (1998). Visceral Perception: Understanding Internal Cognition. New York: Plenum Press.CrossRefGoogle Scholar
Adams, D. J. and Harper, A. A. (1995). Electrophysiological properties of autonomic ganglia. In The Autonomic Nervous System, Vol. 6, Autonomic Ganglia, ed. McLachlan, E. M.. Chur, Switzerland: Harwood Academic Publisher, pp. 153–212.Google Scholar
Ader, A. and Cohen, N. (1993). Psychoneuroendocrinology: conditioning and stress. Annu. Rev. Physiol., 44, 53–85.Google Scholar
Aggleton, J. P. (ed.) (2000). The Amygdala: A Functional Analysis. Oxford: Oxford University Press.Google Scholar
Aicher, S. A., Kurucz, O. S., Reis, D. J. and Milner, T. A. (1995). Nucleus tractus solitarius efferent terminals synapse on neurons in the caudal ventrolateral medulla that project to the rostral ventrolateral medulla. Brain Res., 693, 51–63.CrossRefGoogle ScholarPubMed
Akasu, T. and Koketsu, K. (1986). Muscarinic transmission. In Autonomic and Enteric Ganglia, ed. Karczmar, K., Koketsu, K. and Nishi, S.. New York, London: Plenum Press, pp. 161–180.CrossRefGoogle Scholar
Akasu, T. and Nishimura, T. (1995). Synaptic transmission and function of parasympathetic ganglia. Prog. Neurobiol., 45, 459–522.CrossRefGoogle ScholarPubMed
Akert, K. (1981). Biological Order in Brain Organization. Selected Works of W. R. Hess. Berlin, Heidelberg, New York: Springer-Verlag.CrossRefGoogle Scholar
Akoev, G. N. (1981). Catecholamines, acetylcholine and excitability of mechanoreceptors. Prog. Neurobiol., 15, 269–294.CrossRefGoogle Scholar
Al Chaer, E. D., Lawand, N. B., Westlund, K. N. and Willis, W. D. (1996a). Visceral nociceptive input into the ventral posterolateral nucleus of the thalamus: a new function for the dorsal column pathway. J. Neurophysiol., 76, 2661–2674.CrossRefGoogle Scholar
Al Chaer, E. D., Lawand, N. B., Westlund, K. N. and Willis, W. D. (1996b). Pelvic visceral input into the nucleus gracilis is largely mediated by the postsynaptic dorsal column pathway. J. Neurophysiol., 76, 2675–2690.CrossRefGoogle Scholar
Al Chaer, E. D., Westlund, K. N. and Willis, W. D. (1997). Nucleus gracilis: an integrator for visceral and somatic information. J. Neurophysiol., 78, 521–527.CrossRefGoogle ScholarPubMed
Al Chaer, E. D., Feng, Y. and Willis, W. D. (1999). Comparative study of viscerosomatic input onto postsynaptic dorsal column and spinothalamic tract neurons in the primate. J. Neurophysiol., 82, 1876–1882.CrossRefGoogle ScholarPubMed
Alexander, R. S. (1946). Tonic and reflex functions of medullary sympathetic cardiovascular centers. J. Neurophysiol., 9, 205–217.CrossRefGoogle ScholarPubMed
Alexander, S. P. H., Mathie, A. and Peters, J. A. (2004). Guide to receptors and channels, 1st edition. Br. J. Pharmacol., 141 Suppl1, S1–S126.Google ScholarPubMed
Altschuler, S. M., Bao, X. M., Bieger, D., Hopkins, D. A. and Miselis, R. R. (1989). Viscerotopic representation of the upper alimentary tract in the rat: sensory ganglia and nuclei of the solitary and spinal trigeminal tracts. J. Comp. Neurol., 283, 248–268.CrossRefGoogle ScholarPubMed
Altschuler, S. M., Ferenci, D. A., Lynn, R. B. and Miselis, R. R. (1991). Representation of the cecum in the lateral dorsal motor nucleus of the vagus nerve and commissural subnucleus of the nucleus tractus solitarii in rat. J. Comp. Neurol., 304, 261–274.CrossRefGoogle ScholarPubMed
Altschuler, S. M., Rinaman, L. and Miselis, R. R. (1992). Viscerotopic representation of the alimentary tract in the dorsal and ventral vagal complexes in the rat. In Neuroanatomy and Physiology of Abdominal Vagal Afferents, ed. Ritter, S., Ritter, R. C. and Barnes, C. D.. Boca Raton: CRC Press, pp. 22–53.Google Scholar
Amann, R., Dray, A. and Hankins, M. W. (1988). Stimulation of afferent fibres of the guinea-pig ureter evokes potentials in inferior mesenteric ganglion neurones. J. Physiol., 402, 543–553.CrossRefGoogle ScholarPubMed
Amendt, K., Czachurski, J., Dembowsky, K. and Seller, H. (1979). Bulbospinal projections to the intermediolateral cell column: a neuroanatomical study. J. Auton. Nerv. Syst., 1, 103–107.CrossRefGoogle ScholarPubMed
An, X., Bandler, R., Öngür, D. and Price, J. L. (1998). Prefrontal cortical projections to longitudinal columns in the midbrain periaqueductal gray in macaque monkeys. J. Comp. Neurol., 401, 455–479.3.0.CO;2-6>CrossRefGoogle ScholarPubMed
Anders, S., Lotze, M., Erb, M., Grodd, W. and Birbaumer, N. (2004). Brain activity underlying emotional valence and arousal: a response-related fMRI study. Hum. Brain Mapp., 23, 200–209.CrossRefGoogle ScholarPubMed
Anderson, C. R., McAllen, R. M. and Edwards, S. L. (1995). Nitric oxide synthase and chemical coding in cat sympathetic postganglionic neurons. Neuroscience, 68, 255–264.CrossRefGoogle ScholarPubMed
Anderson, D. J. (1993). Molecular control of cell fate in the neural crest: the sympathoadrenal lineage. Annu. Rev. Neurosci., 16, 129–158.CrossRefGoogle ScholarPubMed
Anderson, E. A., Wallin, B. G. and Mark, A. L. (1987). Dissociation of sympathetic nerve activity in arm and leg muscle during mental stress. Hypertension, 9, III, 114–119.CrossRefGoogle ScholarPubMed
Anderson, R. L., Gibbins, I. L. and Morris, J. L. (1996). Non-noradrenergic sympathetic neurons project to extramuscular feed arteries and proximal intramuscular arteries of skeletal muscles in guinea-pig hindlimbs. J. Auton. Nerv. Syst., 61, 51–60.CrossRefGoogle ScholarPubMed
Andersson, K. E. (2001). Pharmacology of penile erection. Pharmacol. Rev., 53, 417–450.Google ScholarPubMed
Andreev, N. Y., Dimitrieva, N., Koltzenburg, M. and McMahon, S. B. (1995). Peripheral administration of nerve growth factor in the adult rat produces a thermal hyperalgesia that requires the presence of sympathetic post-ganglionic neurones. Pain, 63, 109–115.CrossRefGoogle ScholarPubMed
Andrew, D. and Craig, A. D. (2001a). Spinothalamic lamina I neurones selectively responsive to cutaneous warming in cats. J. Physiol., 537, 489–495.CrossRefGoogle Scholar
Andrew, D. and Craig, A. D. (2001b). Spinothalamic lamina I neurons selectively sensitive to histamine: a central neural pathway for itch. Nat. Neurosci., 4, 72–77.CrossRefGoogle Scholar
Andrew, D. and Craig, A. D. (2002a). Quantitative responses of spinothalamic lamina I neurones to graded mechanical stimulation in the cat. J. Physiol., 545, 913–931.CrossRefGoogle Scholar
Andrew, D. and Craig, A. D. (2002b). Responses of spinothalamic lamina I neurons to maintained noxious mechanical stimulation in the cat. J. Neurophysiol., 87, 1889–1901.CrossRefGoogle Scholar
Andrews, P. L. R. (1986). Vagal afferent innervation of the gastrointestinal tract. Prog. Brain Res., 67, 65–86.CrossRefGoogle ScholarPubMed
Anrep, G. V., Pascual, W. and Rössler, R. (1936a). Respiratory variations of the heart rate. I. The reflex mechanism of the sinus arrhythmia. Proc. Royal Soc. Lond., 119 (Series B), 191–217.CrossRefGoogle Scholar
Anrep, G. V., Pascual, W. and Rössler, R. (1936b). Respiratory variations of the heart rate. II. The central mechanism of the sinus arrhythmia and the inter-relationships between central and reflex mechanism. Proc. Royal Soc. Lond., 119 (Series B), 218–230.CrossRefGoogle Scholar
Apkarian, A. V. and Shi, T. (1994). Squirrel monkey lateral thalamus. I. Somatic nociresponsive neurons and their relation to spinothalamic terminals. J. Neurosci., 14, 6779–6795.CrossRefGoogle ScholarPubMed
Apodaca, G. (2004). The uroepithelium: not just a passive barrier. Traffic, 5, 117–128.CrossRefGoogle Scholar
Apodaca, G., Kiss, S., Ruiz, W.et al. (2003). Disruption of bladder epithelium barrier function after spinal cord injury. Am. J. Physiol. Renal Physiol., 284, F966–F976.CrossRefGoogle ScholarPubMed
Appenzeller, O. and Oribe, E. (1997). The Autonomic Nervous System. An Introduction to Basic and Clinical Concepts, 5th edn. Amsterdam: Elsevier.Google Scholar
Appenzeller, O. (ed.) (1999). Handbook of Clinical Neurology, Vol. 74, The Autonomic Nervous System, part I: Normal Functions. Amsterdam: Elsevier.Google Scholar
Appenzeller, O. (ed.) (2000). Handbook of Clinical Neurology, Vol. 75, The Autonomic Nervous System, part II: Dysfunctions. Amsterdam: Elsevier.Google Scholar
Applegate, C. D., Kapp, B. S., Underwood, M. D. and McNall, C. L. (1983). Autonomic and somatomotor effects of amygdala central N. stimulation in awake rabbits. Physiol. Behav., 31, 353–360.CrossRefGoogle ScholarPubMed
Araki, I. (1994). Inhibitory postsynaptic currents and the effects of γ-aminobutyric acid on visually identified sacral parasympathetic preganglionic neurons in neonatal rats. J. Neurophysiol., 72, 2903–2910.CrossRefGoogle ScholarPubMed
Araki, I. and Groat, W. C. (1996). Unitary excitatory synaptic currents in preganglionic neurons mediated by two distinct groups of interneurons in neonatal rat sacral parasympathetic nucleus. J. Neurophysiol., 76, 215–226.CrossRefGoogle ScholarPubMed
Araki, I. and Groat, W. C. (1997). Developmental synaptic depression underlying reorganization of visceral reflex pathways in the spinal cord. J. Neurosci., 17, 8402–8407.CrossRefGoogle ScholarPubMed
Bachoo, M. and Polosa, C. (1985). Properties of a sympatho-inhibitory and vasodilator reflex evoked by superior laryngeal nerve afferents in the cat. J. Physiol., 364, 183–198.CrossRefGoogle ScholarPubMed
Bachoo, M. and Polosa, C. (1986). The pattern of sympathetic neurone activity during expiration in the cat. J. Physiol., 378, 375–390.CrossRefGoogle ScholarPubMed
Bachoo, M. and Polosa, C. (1987). Properties of the inspiration-related activity of sympathetic preganglionic neurones of the cervical trunk in the cat. J. Physiol., 385, 545–564.CrossRefGoogle ScholarPubMed
Bacon, S. J., Zagon, A. and Smith, A. D. (1990). Electron microscopic evidence of a monosynaptic pathway between cells in the caudal raphe nuclei and sympathetic preganglionic neurons in the rat spinal cord. Exp. Brain Res., 79, 589–602.CrossRefGoogle ScholarPubMed
Bacon, S. J. and Smith, A. D. (1993). A monosynaptic pathway from an identified vasomotor centre in the medial prefrontal cortex to an autonomic area in the thoracic spinal cord. Neuroscience, 54, 719–728.CrossRefGoogle Scholar
Badoer, E. (2001). Hypothalamic paraventricular nucleus and cardiovascular regulation. Clin. Exp. Pharmacol. Physiol., 28, 95–99.CrossRefGoogle ScholarPubMed
Bahns, E., Ernsberger, U., Jänig, W. and Nelke, A. (1986a). Functional characteristics of lumbar visceral afferent fibres from the urinary bladder and the urethra in the cat. Pflügers Arch., 407, 510–518.CrossRefGoogle Scholar
Bahns, E., Ernsberger, U., Jänig, W. and Nelke, A. (1986b). Discharge properties of mechanosensitive afferents supplying the retroperitoneal space. Pflügers Arch., 407, 519–525.CrossRefGoogle Scholar
Bahns, E., Halsband, U. and Jänig, W. (1987). Responses of sacral visceral afferents from the lower urinary tract, colon and anus to mechanical stimulation. Pflügers Arch., 410, 296–303.CrossRefGoogle ScholarPubMed
Bahr, R., Blumberg, H. and Jänig, W. (1981). Do dichotomizing afferent fibers exist which supply visceral organs as well as somatic structures? A contribution to the problem of referred pain. Neurosci. Lett., 24, 25–28.CrossRefGoogle ScholarPubMed
Bahr, R., Bartel, B., Blumberg, H. and Jänig, W. (1986a). Functional characterization of preganglionic neurons projecting in the lumbar splanchnic nerves: neurons regulating motility. J. Auton. Nerv. Syst., 15, 109–130.CrossRefGoogle Scholar
Bahr, R., Bartel, B., Blumberg, H. and Jänig, W. (1986b). Functional characterization of preganglionic neurons projecting in the lumbar splanchnic nerves: vasoconstrictor neurons. J. Auton. Nerv. Syst., 15, 131–140.CrossRefGoogle Scholar
Bahr, R., Bartel, B., Blumberg, H. and Jänig, W. (1986c). Secondary functional properties of lumbar visceral preganglionic neurons. J. Auton. Nerv. Syst., 15, 141–152.CrossRefGoogle Scholar
Bainton, C. R., Richter, D. W., Seller, H., Ballantyne, D. and Klein, J. P. (1985). Respiratory modulation of sympathetic activity. J. Auton. Nerv. Syst., 12, 77–90.CrossRefGoogle ScholarPubMed
Baker, D. G., Coleridge, H. M., Coleridge, J. C. and Nerdrum, T. (1980). Search for a cardiac nociceptor: stimulation by bradykinin of sympathetic afferent nerve endings in the heart of the cat. J. Physiol., 306, 519–536.CrossRefGoogle ScholarPubMed
Baldissera, F., Hultborn, H. and Illert, M. (1981). Integration in spinal neuronal systems. In Handbook of Physiology, Section 1, The Nervous System, Vol. IIb, Motor Control, part I. ed. Brooks, V. B.. Bethesda: American Physiological Society, pp. 509–595.Google Scholar
Baldwin, B. A., Parrott, R. F. and Ebenezer, I. S. (1998). Food for thought: a critique on the hypothesis that endogenous cholecystokinin acts as a physiological satiety factor. Prog. Neurobiol., 55, 477–507.CrossRefGoogle ScholarPubMed
Bamshad, M., Aoki, V. T., Adkison, M. G., Warren, W. S. and Bartness, T. J. (1998). Central nervous system origins of the sympathetic nervous system outflow to white adipose tissue. Am. J. Physiol., 275, R291–R299.Google ScholarPubMed
Bamshad, M., Song, C. K. and Bartness, T. J. (1999). central nervous system origins of the sympathetic nervous system outflow to brown adipose tissue. Am. J. Physiol., 276, R1569–R1578.Google ScholarPubMed
Bandler, R. (1988). Brain mechanisms of aggression as revealed by electrical and chemical stimulation: suggestion of a central role for the midbrain periaqueductal grey region. Prog. Psychobiol. Physiol. Psychol., 13, 67–153.Google Scholar
Bandler, R. and Shipley, M. T. (1994). Columnar organization in the midbrain periaqueductal gray: modules for emotional expression?Trends Neurosci., 17, 379–389.CrossRefGoogle ScholarPubMed
Bandler, R. and Keay, K. A. (1996). Columnar organization in the midbrain periaqueductal gray and the integration of emotional expression. Prog. Brain Res., 107, 285–300.CrossRefGoogle ScholarPubMed
Bandler, R., Carrive, P. and Zhang, S. P. (1991). Integration of somatic and autonomic reactions within the midbrain periaqueductal grey: viscerotopic, somatotopic and functional organization. Prog. Brain Res., 87, 269–305.CrossRefGoogle ScholarPubMed
Bandler, R., Keay, K. A., Floyd, N. and Price, J. (2000a). Central circuits mediating patterned autonomic activity during active vs. passive emotional coping. Brain Res. Bull., 53, 95–104.CrossRefGoogle Scholar
Bandler, R., Price, J. L. and Keay, K. A. (2000b). Brain mediation of active and passive emotional coping. Prog. Brain Res., 122, 333–349.CrossRefGoogle Scholar
Bao, J. X., Gonon, F. and Stjärne, L. (1993). Frequency- and train length-dependent variation in the roles of postjunctional alpha 1- and alpha 2-adrenoceptors for the field stimulation-induced neurogenic contraction of rat tail artery. Naunyn-Schmiedeberg's Arch. Pharmacol., 347, 601–616.CrossRefGoogle ScholarPubMed
Barcroft, H., Brod, Z., Hejl, Z., Hirsjärvi, E. A. and Kitchen, A. H. (1960). The mechanism of the vasodilatation in the forearm muscle during stress (mental arithmetic). Clin. Sci., 19, 577–586.Google Scholar
Bard, P. (1928). A diencephalic mechanism for the expression of rage with special reference to the sympathetic nervous system. Am. J. Physiol., 84, 490–515.Google Scholar
Bard, P. (1932). An emotional expression after decortication with some remarks on certain theoretical views. Part I. Psychol. Rev., 41, 309–329.CrossRefGoogle Scholar
Bard, P. (1960). Anatomical organization of the central nervous system in relation to control of the heart and blood vessels. Physiol. Rev., 40 (Suppl4), 3–26.Google Scholar
Bard, P. and Macht, M. B. (1957). The behavior of chronically decerebrate cats. In Ciba Foundation Symposium on the Neurological Basis of Behavior, ed. Wolstenholme, G. E. W. and O'Connor, M.. Boston: Little, Brown and Co., pp. 55–71.Google Scholar
Bard, P. and Rioch, D. McK. (1937). A study of four cats deprived of neocortex and additional portions of the forebrain. Bull. Johns Hopkins Hosp., 60, 73–147.Google Scholar
Barker, D. and Saito, M. (1981). Autonomic innervation of receptors and muscle fibres in cat skeletal muscle. Proc. R. Soc. London Biol., 212, 317–332.CrossRefGoogle ScholarPubMed
Barman, S. M. and Gebber, G. L. (1976). Basis for synchronization of sympathetic and phrenic nerve discharges. Am. J. Physiol., 231, 1601–1607.Google ScholarPubMed
Barman, S. M. and Gebber, G. L. (1984). Spinal interneurons with sympathetic nerve-related activity. Am. J. Physiol., 247, R761–R767.Google ScholarPubMed
Barman, S. M. and Gebber, G. L. (2000). “Rapid” rhythmic discharges of sympathetic nerves: sources, mechanisms of generation, and physiological relevance. J. Biol. Rhythms, 15, 365–379.CrossRefGoogle ScholarPubMed
Barman, S. M., Orer, H. S. and Gebber, G. L. (2001). The role of the medullary lateral tegmental field in the generation and baroreceptor reflex control of sympathetic nerve discharge in the cat. Ann. N. Y. Acad. Sci., 940, 270–285.CrossRefGoogle ScholarPubMed
Barnes, P. J. (ed.) (1997). The Autonomic Nervous System, Vol. 7, Autonomic Control of the Respiratory System. Amsterdam: Harwood Academic Publishers.Google Scholar
Baron, R. and Jänig, W. (1988). Neurons projecting rostrally in the hypogastric nerve of the cat. J. Auton. Nerv. Syst., 24, 81–86.CrossRefGoogle ScholarPubMed
Baron, R. and Jänig, W. (1991). Afferent and sympathetic neurons projecting into lumbar visceral nerves of the male rat. J. Comp. Neurol., 314, 429–436.CrossRefGoogle ScholarPubMed
Baron, R., Jänig, W. and McLachlan, E. M. (1985a). On the anatomical organization of the lumbosacral sympathetic chain and the lumbar splanchnic nerves of the cat – Langley revisited. J. Auton. Nerv. Syst., 12, 289–300.CrossRefGoogle Scholar
Baron, R., Jänig, W. and McLachlan, E. M. (1985b). The afferent and sympathetic components of the lumbar spinal outflow to the colon and pelvic organs in the cat: I. The hypogastric nerve. J. Comp. Neurol., 238, 135–146.CrossRefGoogle Scholar
Baron, R., Jänig, W. and McLachlan, E. M. (1985c). The afferent and sympathetic components of the lumbar spinal outflow to the colon and pelvic organs in the cat. II. The lumbar splanchnic nerves. J. Comp. Neurol., 238, 147–157.CrossRefGoogle Scholar
Baron, R., Jänig, W. and McLachlan, E. M. (1985d). The afferent and sympathetic components of the lumbar spinal outflow to the colon and pelvic organs in the cat. III. The colonic nerves, incorporating an analysis of all components of the lumbar prevertebral outflow. J. Comp. Neurol., 238, 158–168.CrossRefGoogle Scholar
Baron, R., Jänig, W. and Kollmann, W. (1988). Sympathetic and afferent somata projecting in hindlimb nerves and the anatomical organization of the lumbar sympathetic nervous system of the rat. J. Comp. Neurol., 275, 460–468.CrossRefGoogle ScholarPubMed
Baron, R., Jänig, W. and With, H. (1995). Sympathetic and afferent neurones projecting into forelimb and trunk nerves and the anatomical organization of the thoracic sympathetic outflow of the rat. J. Auton. Nerv. Syst., 53, 205–214.CrossRefGoogle ScholarPubMed
Barraco, I. R. A. (ed.) (1994). Nucleus of the Solitary Tract. Boca Raton: CRC Press.Google Scholar
Bartel, B., Blumberg, H. and Jänig, W. (1986). Discharge patterns of motility-regulating neurons projecting in the lumbar splanchnic nerves to visceral stimuli in spinal cats. J. Auton. Nerv. Syst., 15, 153–163.CrossRefGoogle ScholarPubMed
Bartness, T. J. and Bamshad, M. (1998). Innervation of mammalian white adipose tissue: implications for the regulation of total body fat. Am. J. Physiol., 275, R1399–R1411.Google ScholarPubMed
Bartsch, T., Häbler, H. J. and Jänig, W. (1996). Functional properties of postganglionic sympathetic neurones supplying the submandibular gland in the anaesthetized rat. Neurosci. Lett., 214, 143–146.CrossRefGoogle ScholarPubMed
Bartsch, T., Häbler, H. J. and Jänig, W. (1999). Hypoventilation recruits preganglionic sympathetic fibers with inspiration-related activity in the superior cervical trunk of the rat. J. Auton. Nerv. Syst., 77, 31–38.CrossRefGoogle ScholarPubMed
Bartsch, T., Häbler, H. J. and Jänig, W. (2000). Reflex patterns of preganglionic sympathetic fibers projecting to the superior cervical ganglion in the rat. Auton. Neurosci., 83, 66–74.CrossRefGoogle ScholarPubMed
Baumann, T. K., Simone, D. A., Shain, C. N. and LaMotte, R. H. (1991). Neurogenic hyperalgesia: the search for the primary cutaneous afferent fibers that contribute to capsaicin-induced pain and hyperalgesia. J. Neurophysiol., 66, 212–227.CrossRefGoogle ScholarPubMed
Bayliss, W. M. (1901). On the origin from the spinal cord of the vaso-dilator fibres of the hind-limb, and on the nature of these fibres. J. Physiol., 26, 173–209.CrossRefGoogle ScholarPubMed
Bayliss, W. M. and Starling, E. H. (1899). The movements and innervation of the small intestine. J. Physiol., 24, 99–143.CrossRefGoogle ScholarPubMed
Bayliss, W. M. and Starling, E. H. (1900). The movements and innervation of the large intestine. J. Physiol., 26, 107–118.CrossRefGoogle ScholarPubMed
Beckstead, R. M., Morse, J. R. and Norgren, R. (1980). The nucleus of the solitary tract in the monkey: projections to the thalamus and brain stem nuclei. J. Comp. Neurol., 190, 259–282.CrossRefGoogle ScholarPubMed
Bell, C., Jänig, W., Kümmel, H. and Xu, H. (1985). Differentiation of vasodilator and sudomotor responses in the cat paw pad to preganglionic sympathetic stimulation. J. Physiol., 364, 93–104.CrossRefGoogle ScholarPubMed
Belmonte, C. and Cervero, F. (eds.) (1996). Neurobiology of Nociception. Oxford, New York, Tokyo: Oxford University Press.CrossRefGoogle Scholar
Benison, S., Barger, A. C. and Wolfe, E. L. (1987). Walter B. Cannon. The Life and Times of a Young Scientist. Cambridge Mass. London England: The Belknarp Press of Harvard University Press.Google Scholar
Bennett, T. and Gardiner, S. M. (eds.) (1996). The Autonomic Nervous System, Vol. 8, Nervous Control of Blood Vessels. Amsterdam: Harwood Academic Publishers.Google Scholar
Berkley, K. J., Robbins, A. and Sato, Y. (1988). Afferent fibers supplying the uterus in the rat. J. Neurophysiol., 59, 142–163.CrossRefGoogle ScholarPubMed
Berkley, K. J., Hotta, H., Robbins, A. and Sato, Y. (1990). Functional properties of afferent fibers supplying reproductive and other pelvic organs in pelvic nerve of female rat. J. Neurophysiol., 63, 256–272.CrossRefGoogle ScholarPubMed
Berkley, K. J., Robbins, A. and Sato, Y. (1993). Functional differences between afferent fibers in the hypogastric and pelvic nerves innervating female reproductive organs in the rat. J. Neurophysiol., 69, 533–544.CrossRefGoogle ScholarPubMed
Bernard, C. (1858). Leçons Sur la Physiologie et la Pathologie du Système Nerveau [On the Physiology and Pathophysiology of the Nervous System]. Paris: Bailliere.Google Scholar
Bernard, C. (1957) [1865]. Introduction à l'Étude de la Médicine Experimentale. [An Introduction to the Study of Experimental Medicine]. New York: Dover Publications. Paris: Baillière.Google Scholar
Bernard, C. (1974) [1878]. Lecons sur les Phénomènes de la Vie Communes aux Animaux et aux Végétaux [Lectures on the Phenomena of Life Common to Animals and Plants]. Translated by H. E. Hoff, R. Guillemin and L. Guillemin. Paris (Springfield Illinois): B. Ballière et Fils (Thomas).Google Scholar
Bernard, J. F. and Bandler, R. (1998). Parallel circuits for emotional coping behaviour: new pieces in the puzzle. J. Comp. Neurol., 401, 429–436.3.0.CO;2-3>CrossRefGoogle ScholarPubMed
Bernard, J. F. and Besson, J. M. (1990). The spino(trigemino)pontoamygdaloid pathway: electrophysiological evidence for an involvement in pain processes. J. Neurophysiol., 63, 473–490.CrossRefGoogle ScholarPubMed
Bernard, J. F., Huang, G. F. and Besson, J. M. (1994). The parabrachial area: electrophysiological evidence for an involvement in visceral nociceptive processes. J. Neurophysiol., 71, 1646–1660.CrossRefGoogle ScholarPubMed
Berne, C. and Fagius, J. (1986). Skin sympathetic activity during insulin-induced hypoglycemia. Diabetologia, 29, 855–860.CrossRefGoogle Scholar
Berntson, G. G. and Cacioppo, J. T. (2000). From homeostasis to allodynamic regulation. In Handbook of Psychophysiology, 2nd edn., eds. Cacioppo, J. T., Tassinary, L. G. and Berntson, G. G.. Cambridge: Cambridge University Press, pp. 459–481.Google Scholar
Berthoud, H. R. and Neuhuber, W. L. (2000). Functional and chemical anatomy of the afferent vagal system. Auton. Neurosci., 85, 1–17.CrossRefGoogle ScholarPubMed
Berthoud, H. R. and Powley, T. L. (1990). Identification of vagal preganglionics that mediate cephalic phase insulin response. Am. J. Physiol., 258, R523–R530.Google ScholarPubMed
Berthoud, H. R. and Powley, T. L. (1992). Vagal afferent innervation of the rat fundic stomach: morphological characterization of the gastric tension receptor. J. Comp. Neurol., 319, 261–276.CrossRefGoogle ScholarPubMed
Berthoud, H. R., Fox, E. A. and Powley, T. L. (1990). Localization of vagal preganglionics that stimulate insulin and glucagon secretion. Am. J. Physiol., 258, R160–R168.Google ScholarPubMed
Berthoud, H. R., Carlson, N. R. and Powley, T. L. (1991). Topography of efferent vagal innervation of the rat gastrointestinal tract. Am. J. Physiol., 260, R200–R207.Google ScholarPubMed
Berthoud, H. R., Patterson, L. M., Willing, A. E., Mueller, K. and Neuhuber, W. L. (1997). Capsaicin-resistant vagal afferent fibers in the rat gastrointestinal tract: anatomical identification and functional integrity. Brain Res., 746, 195–206.CrossRefGoogle ScholarPubMed
Besedovsky, H. O. and del Rey, A. (1992). Immune-neuroendocrine circuits: integrative role of cytokines. Front. Neuroendocrinol., 13, 61–94.Google ScholarPubMed
Besedovsky, H. O. and del Rey, A. (1995). Immune-neuroendocrine interactions: facts and hypotheses. Endocr. Rev., 17, 64–102.CrossRefGoogle Scholar
Bester, H., Chapman, V., Besson, J. M. and Bernard, J. F. (2000). Physiological properties of the lamina I spinoparabrachial neurons in the rat. J. Neurophysiol., 83, 2239–2259.CrossRefGoogle ScholarPubMed
Beyak, M. J. and Grundy, D. (2005). Vagal afferents innervating the gastrointestinal tract. In Advance in Vagal Afferent Neurobiology, ed. Undem, B. and Weinreich, D.. Boca Raton: CRC Press, pp. 315–350.CrossRefGoogle Scholar
Bianchi, A. L., Denavit-Saubie, M. and Champagnat, J. (1995). Central control of breathing in mammals: neuronal circuitry, membrane properties, and neurotransmitters. Physiol. Rev., 75, 1–45.CrossRefGoogle ScholarPubMed
Bieger, D. and Hopkins, D. A. (1987). Viscerotopic representation of the upper alimentary tract in the medulla oblongata in the rat: the nucleus ambiguus. J. Comp. Neurol., 262, 546–562.CrossRefGoogle ScholarPubMed
Bielefeld, T. K. and Gebhart, G. F. (2005). Visceral pain: basic mechanisms. In Wall and Melzack's Textbook of Pain, 5th edn., ed. McMahon, S. B. and Koltzenburg, M.. Amsterdam, Edinburgh: Elsevier Churchill Livingstone, pp. 721–736.Google Scholar
Bini, G., Hagbarth, K. E., Hynninen, P. and Wallin, B. G. (1980a). Thermoregulatory and rhythm-generating mechanisms governing the sudomotor and vasoconstrictor outflow in human cutaneous nerves. J. Physiol., 306, 547–552.CrossRefGoogle Scholar
Bini, G., Hagbarth, K. E., Hynninen, P. and Wallin, B. G. (1980b). Regional similarities and differences in thermoregulatory vaso- and sudomotor tone. J. Physiol., 306, 553–565.CrossRefGoogle Scholar
Bini, G., Hagbarth, K. E. and Wallin, B. G. (1981). Cardiac rhythmicity of skin sympathetic activity recorded from peripheral nerves in man. J. Auton. Nerv. Syst., 4, 17–24.CrossRefGoogle ScholarPubMed
Birder, L. A. (2005). More than just a barrier: urothelium as a drug target for urinary bladder pain. Am. J. Physiol. Renal. Physiol., 289, F489–F495.CrossRefGoogle ScholarPubMed
Björntop, P. (1997). Behavior and metabolic disease. Int. J. Behav. Med., 3, 285–302.CrossRefGoogle Scholar
Blackman, J. G. (1974). Function of autonomic ganglia. In The Peripheral Nervous System, ed. Hubbard, J. I.. New York: Plenum Press, pp. 257–276.CrossRefGoogle Scholar
Blackman, J. G., Ginsborg, B. L. and Ray, C. (1963). Synaptic transmission in the sympathetic ganglion of the frog. J. Physiol., 167, 355–373.CrossRefGoogle ScholarPubMed
Blair, D. A., Glover, W. E., Greenfield, A. D. M. and Roddie, I. C. (1959). Excitation of cholinergic vasodilator nerves to human skeletal muscles during emotional stress. J. Physiol., 148, 633–646.CrossRefGoogle ScholarPubMed
Blessing, W. W. (1997). The Lower Brain Stem and Bodily Homeostasis. New York, Oxford: Oxford University Press.Google Scholar
Blessing, W. W. and Nalivaiko, E. (2000). Regional blood flow and nociceptive stimuli in rabbits: patterning by medullary raphe, not ventrolateral medulla. J. Physiol., 524, 279–292.CrossRefGoogle Scholar
Blessing, W. W., Li, Y. W. and Wesselingh, S. L. (1991). Transneuronal transport of herpes simplex virus from the cervical vagus to brain neurons with axonal inputs to central vagal sensory nuclei in the rat. Neuroscience, 42, 261–274.CrossRefGoogle ScholarPubMed
Blessing, W. W., Yu, Y. H. and Nalivaiko, E. (1999). Raphe pallidus and parapyramidal neurons regulate ear pinna vascular conductance in the rabbit. Neurosci. Lett., 270, 33–36.CrossRefGoogle ScholarPubMed
Blix, A. S. and Folkow, B. (1983). Cardiovascular adjustments to diving in mammals and birds. In Handbook of Physiology, Section 2: The Cardiovascular System Vol. III: Peripheral Circulation, ed. Shepherd, J. T. and Abboud, F. M.. Bethesda: American Physiological Society, pp. 917–945.Google Scholar
Blok, B. F. and Holstege, G. (1996). The neuronal control of micturition and its relation to the emotional motor system. Prog. Brain Res., 107, 113–126.CrossRefGoogle ScholarPubMed
Blumberg, H. and Jänig, W. (1982). Changes in unmyelinated fibers including sympathetic postganglionic fibers of a skin nerve after peripheral neuroma formation. J. Auton. Nerv. Syst., 6, 173–183.CrossRefGoogle ScholarPubMed
Blumberg, H. and Jänig, W. (1983a). Enhancement of resting activity in postganglionic vasoconstrictor neurones following short-lasting repetitive activation of preganglionic axons. Pflügers Arch., 396, 89–94.CrossRefGoogle Scholar
Blumberg, H. and Jänig, W. (1983b). Changes of reflexes in vasoconstrictor neurons supplying the cat hindlimb following chronic nerve lesions: a model for studying mechanisms of reflex sympathetic dystrophy? J. Auton. Nerv. Syst., 7, 399–411.CrossRefGoogle Scholar
Blumberg, H. and Jänig, W. (1985). Reflex patterns in postganglionic vasoconstrictor neurons following chronic nerve lesions. J. Auton. Nerv. Syst., 14, 157–180.CrossRefGoogle ScholarPubMed
Blumberg, H. and Wallin, B. G. (1987). Direct evidence of neurally mediated vasodilatation in hairy skin of the human foot. J. Physiol., 382, 105–121.CrossRefGoogle ScholarPubMed
Blumberg, H., Jänig, W., Rieckmann, C. and Szulczyk, P. (1980). Baroreceptor and chemoreceptor reflexes in postganglionic neurones supplying skeletal muscle and hairy skin. J. Auton. Nerv. Syst., 2, 223–240.CrossRefGoogle ScholarPubMed
Blumberg, H., Haupt, P., Jänig, W. and Kohler, W. (1983). Encoding of visceral noxious stimuli in the discharge patterns of visceral afferent fibres from the colon. Pflügers Arch., 398, 33–40.CrossRefGoogle ScholarPubMed
Boczek-Funcke, A., Häbler, H. J., Jänig, W. and Michaelis, M. (1991). Rapid phasic baroreceptor inhibition of the activity in sympathetic preganglionic neurones does not change throughout the respiratory cycle. J. Auton. Nerv. Syst., 34, 185–194.CrossRefGoogle Scholar
Boczek-Funcke, A., Dembowsky, K., Häbler, H. J.et al. (1992a). Classification of preganglionic neurones projecting into the cat cervical sympathetic trunk. J. Physiol., 453, 319–339.CrossRefGoogle Scholar
Boczek-Funcke, A., Dembowsky, K., Häbler, H. J., Jänig, W. and Michaelis, M. (1992b). Respiratory-related activity patterns in preganglionic neurones projecting into the cat cervical sympathetic trunk. J. Physiol., 457, 277–296.CrossRefGoogle Scholar
Boczek-Funcke, A., Häbler, H. J., Jänig, W. and Michaelis, M. (1992c). Respiratory modulation of the activity in sympathetic neurones supplying muscle, skin and pelvic organs in the cat. J. Physiol., 449, 333–361.CrossRefGoogle Scholar
Boczek-Funcke, A., Dembowsky, K., Häbler, H. J., Jänig, W. and Michaelis, M. (1993). Spontaneous activity, conduction velocity and segmental origin of different classes of thoracic preganglionic neurons projecting into the cat cervical sympathetic trunk. J. Auton. Nerv. Syst., 43, 189–200.CrossRefGoogle ScholarPubMed
Bogduk, N. (1983). The innervation of the lumbar spine. Spine, 8, 286–293.CrossRefGoogle ScholarPubMed
Bogduk, N., Windsor, M. and Inglis, A. (1988). The innervation of the cervical intervertebral discs. Spine, 13, 2–8.CrossRefGoogle ScholarPubMed
Bolme, P. and Fuxe, K. (1970). Adrenergic and cholinergic nerve terminals in skeletal muscle vessels. Acta Physiol. Scand., 78, 52–59.CrossRefGoogle ScholarPubMed
Bolme, B., Novotny, J., Uvnäs, B. and Wright, P. G. (1970). Species distribution of sympathetic cholinergic vasodilator nerves in skeletal muscle. Acta Physiol. Scand., 78, 60–64.CrossRefGoogle ScholarPubMed
Bolter, C. P., Wallace, D. J. and Hirst, G. D. (2001). Failure of Ba2 + and Cs+ to block the effects of vagal nerve stimulation in sinoatrial node cells of the guinea-pig heart. Auton. Neurosci., 94, 93–101.CrossRefGoogle ScholarPubMed
Bolton, B., Carmichael, E. A. and Stürup, G. (1936). Vaso-constriction following deep inspiration. J. Physiol., 86, 83–94.CrossRefGoogle ScholarPubMed
Bors, E. H. and Comarr, A. E. (1960). Neurological disturbances of sexual function with special reference to 529 patients with spinal cord injury. Urol. Survey, 10, 191–222.Google Scholar
Bos, J. D. (ed.) (1989). Skin Immune System. Boca Raton: CRC Press.Google Scholar
Bos, J. D. and Kapsenberg, M. L. (1986). The skin immune system. Its cellular constituents and their interactions. Immunol. Today, 7, 235–240.CrossRefGoogle ScholarPubMed
Boscan, P. and Paton, J. F. (2002). Integration of cornea and cardiorespiratory afferents in the nucleus of the solitary tract of the rat. Am. J. Physiol. Heart Circ. Physiol., 282, H1278–H1287.CrossRefGoogle ScholarPubMed
Boscan, P., Kasparov, S. and Paton, J. F. (2002). Somatic nociception activates neurokinin 1 receptors in the nucleus tractus solitarii to attenuate the baroreceptor cardiac reflex. Eur. J. Neurosci., 16, 907–920.CrossRefGoogle ScholarPubMed
Bosnjak, Z. J. and Kampine, J. P. (1982). Intracellular recordings from the stellate ganglion of the cat. J. Physiol., 324, 273–283.CrossRefGoogle ScholarPubMed
Bosnjak, Z. J. and Kampine, J. P. (1984). Peripheral neural input to neurons of the middle cervical ganglion in the cat. Am. J. Physiol., 246, R354–R358.Google ScholarPubMed
Bosnjak, Z. J. and Kampine, J. P. (1985). Electrophysiological and morphological characterization of neurons in stellate ganglion of cats. Am. J. Physiol., 248, R288–R292.Google ScholarPubMed
Boucsein, W. (1992). Electrodermal Activity. New York: Plenum Press.CrossRefGoogle Scholar
Boyd, H. D., McLachlan, E. M., Keast, J. R. and Inokuchi, H. (1996). Three electrophysiological classes of guinea pig sympathetic postganglionic neurone have distinct morphologies. J. Comp. Neurol., 369, 372–387.3.0.CO;2-2>CrossRefGoogle ScholarPubMed
Bramich, N. J., Edwards, F. R. and Hirst, G. D. (1990). Sympathetic nerve stimulation and applied transmitters on the sinus venosus of the toad. J. Physiol., 429, 349–375.CrossRefGoogle ScholarPubMed
Bramich, N. J., Brock, J. A., Edwards, F. R. and Hirst, G. D. (1993). Responses to sympathetic nerve stimulation of the sinus venosus of the toad. J. Physiol., 461, 403–430.CrossRefGoogle ScholarPubMed
Bramich, N. J., Brock, J. A., Edwards, F. R. and Hirst, G. D. (1994). Ionophoretically applied acetylcholine and vagal stimulation in the arrested sinus venosus of the toad, Bufo marinus. J. Physiol., 478, 289–300.Google ScholarPubMed
Bramich, N. J., Cousins, H. M., Edwards, F. R. and Hirst, G. D. (2001). Parallel metabotropic pathways in the heart of the toad, Bufo marinus. Am. J. Physiol. Heart Circ. Physiol., 281, H1771–H1777.CrossRefGoogle ScholarPubMed
Brock, J. A. and Cunnane, T. C. (1988). Electrical activity at the sympathetic neuroeffector junction in the guinea-pig vas deferens. J. Physiol., 399, 607–632.CrossRefGoogle ScholarPubMed
Brock, J. A. and Cunnane, T. C. (1992). Impulse conduction in sympathetic nerve terminals in the guinea-pig vas deferens and the role of the pelvic ganglia. Neuroscience, 47, 185–196.CrossRefGoogle ScholarPubMed
Brock, J. A. and Cunnane, T. C. (1993). Neurotransmitter release mechanisms at the sympathetic neuroeffector junction. Exp. Physiol., 78, 591–614.CrossRefGoogle ScholarPubMed
Brock, J. A. and Helden, D. F. (1995). Enhanced excitatory junction potentials in mesenteric arteries from spontaneously hypertensive rats. Pflügers Arch., 430, 901–908.CrossRefGoogle ScholarPubMed
Brock, J. A., McLachlan, E. M. and Rayner, S. E. (1997). Contribution of alpha-adrenoceptors to depolarization and contraction evoked by continuous asynchronous sympathetic nerve activity in rat tail artery. Br. J. Pharmacol., 120, 1513–1521.CrossRefGoogle ScholarPubMed
Brodal, P. (1998). The Central Nervous System. Structure and Function. New York, Oxford: Oxford University Press.Google Scholar
Brooke, R. E., Pyner, S., McLeish, P.et al. (2002). Spinal cord interneurones labelled transneuronally from the adrenal gland by a GFP-herpes virus construct contain the potassium channel subunit Kv3.1b. Auton. Neurosci., 98, 45–50.CrossRefGoogle ScholarPubMed
Brooke, R. E., Deuchars, J. and Deuchars, S. A. (2004). Input-specific modulation of neurotransmitter release in the lateral horn of the spinal cord via adenosine receptors. J. Neurosci., 24, 127–137.CrossRefGoogle ScholarPubMed
Brookes, S. J. (2001). Classes of enteric nerve cells in the guinea-pig small intestine. Anat. Rec., 262, 58–70.3.0.CO;2-V>CrossRefGoogle ScholarPubMed
Brookes, S. and Costa, M. (eds.) (2002). The Autonomic Nervous System, Vol. 14, Innervation of the Gastrointestinal Tract. London, New York: Francis and Taylor.Google Scholar
Brooks, C. M., Koizumi, K. and Pinkston, J. O. (eds.) (1975). The Life and Contributions of Walter Bradford Cannon 1871–1945. New York: State University of New York, Downstate Medical Center.Google Scholar
Brown, A. M. (1967). Cardiac sympathetic adrenergic pathways in which synaptic transmission is blocked by atropine sulfate. J. Physiol., 191, 271–288.CrossRefGoogle ScholarPubMed
Brown, A. M. (1969). Sympathetic ganglionic transmission and the cardiovascular changes of the defense reaction in the cat. Circ. Res., 24, 843–849.CrossRefGoogle ScholarPubMed
Brown, A., Ricci, M. J. and Weaver, L. C. (2004). nerve growth factor message and protein distribution in the injured rat spinal cord. Exp. Neurol., 188, 115–127.CrossRefGoogle ScholarPubMed
Brown, D. L. and Guyenet, P. G. (1984). Cardiovascular neurons of brain stem with projections to spinal cord. Am. J. Physiol., 247, R1009–R1016.Google ScholarPubMed
Brown, D. L. and Guyenet, P. G. (1985). Electrophysiological study of cardiovascular neurons in the rostral ventrolateral medulla in rats. Circ. Res., 56, 359–369.CrossRefGoogle ScholarPubMed
Browning, K. N., Renehan, W. E. and Travagli, R. A. (1999). Electrophysiological and morphological heterogeneity of rat dorsal vagal neurones which project to specific areas of the gastrointestinal tract. J. Physiol., 517, 521–532.CrossRefGoogle ScholarPubMed
Browning, K. N., Coleman, F. H. and Travagli, R. (2005). Effects of pancreatic polypeptide on pancreas-projecting rat dorsal motor nucleus of the vagus neuron. Am. J. Physiol. Gastrointest. Liver Physiol., 289, G209–G219.CrossRefGoogle Scholar
Bruce, A. N. (1910). Über die Beziehung der sensiblen Nervenendigungen zum Entzündungsvorgang [On the relation between sensory nerve endings and inflammation]. Arch. Exptl. Pathol. Pharmakol., 63, 424–433.CrossRefGoogle Scholar
Bruce, A. N. (1913). Vaso-dilator axon-reflexes. Q. J. Exp. Physiol., 6, 339–354.CrossRefGoogle Scholar
Burnett, A. L., Lowenstein, C. J., Bredt, D. S., Chang, T. S. and Snyder, S. H. (1992). Nitric oxide: a physiologic mediator of penile erection. Science, 257, 401–403.CrossRefGoogle ScholarPubMed
Burns, A. J., Lomax, A. E., Torihashi, S., Sanders, K. M. and Ward, S. M. (1996). Interstitial cells of Cajal mediate inhibitory neurotransmission in the stomach. Proc. Natl. Acad. Sci. U.S.A., 93, 12008–12013.CrossRefGoogle ScholarPubMed
Burnstock, G. and Hoyle, C. H. V. (eds.) (1992). The Autonomic Nervous System, Vol. 1, Autonomic Neuroeffector Mechanisms. Chur, Switzerland: Harwood Academic Publishers.Google Scholar
Burnstock, G. and Sillito, A. M. (eds.) (2000). The Autonomic Nervous System, Vol. 13, Nervous Control of the Eye. Amsterdam: Harwood Academic Publishers.Google Scholar
Busse, R., Edwards, G., Feletou, M.et al. (2002). EDHF: bringing the concepts together. Trends Pharmacol. Sci., 23, 374–380.CrossRefGoogle ScholarPubMed
Butler, P. J. and Jones, D. R. (1997). Physiology of diving of birds and mammals. Physiol. Rev., 77, 837–899.CrossRefGoogle ScholarPubMed
Bykov, K. M. (1959) [1944]. The Cerebral Cortex and the Internal Organs. [Translated from Russian and edited by R. Hodes and A. Kilbey.] Moscow: Foreign Language Publishing House.Google Scholar
Cabot, J. B. (1990). Sympathetic preganglionic neurons: cytoarchitecture, ultrastructure, and biophysical properties. In Central Regulation of Autonomic Functions, eds. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 44–67.Google Scholar
Cabot, J. B. (1996). Some principles of the spinal organization of the sympathetic preganglionic outflow. Prog. Brain Res., 107, 29–42.CrossRefGoogle ScholarPubMed
Cabot, J. B., Alessi, V., Carroll, J. and Ligorio, M. (1994). Spinal cord lamina V and lamina VII interneuronal projections to sympathetic preganglionic neurons. J. Comp. Neurol., 347, 515–530.CrossRefGoogle ScholarPubMed
Cajal, S. R. (1995) [1911]. Histologie du Système Nerveux de l'Homme et des Vertèbrès. [Histology of the Nervous System of Man and Vertebrates]. Maloine, Vol. 2, edited and translated by Swanson, L. W. and Swanson, N.. Oxford: Oxford University Press.Google Scholar
Campbell, G. D., Edwards, F. R., Hirst, G. D. S. and O'Shea, J. E. (1989). Effects of vagal stimulation and applied acetylcholine on pacemaker potentials in the guinea-pig heart. J. Physiol., 415, 57–68.CrossRefGoogle ScholarPubMed
Campos, R. R. and McAllen, R. M. (1997). Cardiac sympathetic premotor neurons. Am. J. Physiol., 272, R615–R620.Google ScholarPubMed
Campos, R. R. and McAllen, R. M. (1999). Tonic drive to sympathetic premotor neurons of rostral ventrolateral medulla from caudal pressor area neurons. Am. J. Physiol., 276, R1209–R1213.Google ScholarPubMed
Canning, B. J. and Mazzone, S. B. (2005). Reflexes initiated by activation of the vagal afferent nerves innervating the airways and lung. In Advances in Vagal Afferent Neurobiology, ed. Undem, B. J. and Weinreich, D., Boca Raton: CRC, Taylor & Francis, pp. 403–430.CrossRefGoogle Scholar
Cannon, W. B. (1911). The Mechanical Factors of Digestion. London: Edward Arnold.Google Scholar
Cannon, W. B. (1914a). The interrelations of emotions as suggested by recent physiological researches. Am. J. Physiol., 25, 252–282.Google Scholar
Cannon, W. B. (1914b). The emergency function of the adrenal medulla in pain and the major emotions. Am. J. Physiol., 33, 356–372.Google Scholar
Cannon, W. B. (1927). The James-Lange theory of emotions: a critical examination and an alternative theory. Am. J. Psychol., 39, 106–124.CrossRefGoogle Scholar
Cannon, W. B. (1928). Die Notfallfunktion des sympathico-adrenalen Systems [The emergency function of the sympathico-adrenal system]. Ergebn. Physiol., 27, 380–406.CrossRefGoogle Scholar
Cannon, W. B. (1929a). Organization for physiological homeostasis. Physiol. Rev., 9, 399–431.CrossRefGoogle Scholar
Cannon, W. B. (1929b). Bodily Changes in Pain, Hunger, Fear and Rage. New York: Appleton.Google Scholar
Cannon, W. B. (1933). A method of stimulating autonomic nerves in the anesthetized cat with observation on the motor and sensory effects. Am. J. Physiol., 105, 366–372.Google Scholar
Cannon, W. B. (1939). The Wisdom of the Body, 2nd revised and enlarged edition. New York: Norton.Google Scholar
Cannon, W. B. and Murphy, F. T. (1906). The movements of the stomach and intestine in some surgical conditions. Ann. Surg., 43, 512–536.CrossRefGoogle ScholarPubMed
Cannon, W. B., Newton, H. F., Bright, E. M., Menkin, V. and Moore, R. M. (1929). Some aspects of the physiology of animals surviving complete exclusion of sympathetic nerve impulses. Am. J. Physiol., 89, 84–107.Google Scholar
Cano, G., Card, J. P., Rinaman, L. and Sved, A. F. (2000). Connections of Barrington's nucleus to the sympathetic nervous system in rats. J. Auton. Nerv. Syst., 79, 117–128.CrossRefGoogle ScholarPubMed
Cano, G., Sved, A. F., Rinaman, L., Rabin, B. S. and Card, J. P. (2001). Characterization of the central nervous system innervation of the rat spleen using viral transneuronal tracing. J. Comp. Neurol., 439, 1–18.CrossRefGoogle ScholarPubMed
Card, J. P., Rinaman, L., Schwaber, J. S.et al. (1990). Neurotropic properties of pseudorabies virus: uptake and transneuronal passage in the rat central nervous system. J. Neurosci., 10, 1974–1994.CrossRefGoogle ScholarPubMed
Card, J. P., Swanson, L. W. and Moore, R. Y. (2003). The hypothalamus: an overview of regulatory systems. In Fundamental Neuroscience, 2nd edn., eds. Squire, L. R., Bloom, F. E., McConnell, S. K., et al. San Diego: Academic Press, pp. 897–909.Google Scholar
Carrive, P. (1993). The periaqueductal gray and defensive behavior: functional representation and neuronal organization. Behav. Brain Res., 58, 27–47.CrossRefGoogle ScholarPubMed
Carrive, P. and Morgan, M. M. (2004). Periaqueductal grey. In The Human Nervous System, ed. Paxinos, G. and Mai, J. K.. Amsterdam: Elsevier Academic Press, pp. 393–423.Google Scholar
Cassell, J. F. and McLachlan, E. M. (1986). The effect of a transient outward current (IA) on synaptic potentials in sympathetic ganglion cells of the guinea-pig. J. Physiol., 374, 273–288.CrossRefGoogle ScholarPubMed
Cassell, J. F., Clark, A. L. and McLachlan, E. M. (1986). Characteristics of phasic and tonic sympathetic ganglion cells of the guinea-pig. J. Physiol., 372, 457–483.CrossRefGoogle ScholarPubMed
Cassell, J. F., McLachlan, E. M. and Sittiracha, T. (1988). The effect of temperature on neuromuscular transmission in the main caudal artery of the rat. J. Physiol., 397, 31–49.CrossRefGoogle ScholarPubMed
Casson, D. M. and Ronald, K. (1975). The harp seal, Pagophilus groenlandicus (Erxleben, 1777). XIV. Cardiac arrythmias. Comp. Biochem. Physiol. A, 50, 307–314.CrossRefGoogle ScholarPubMed
Causing, C. G., Gloster, A., Aloyz, R.et al. (1997). Synaptic innervation density is regulated by neuron-derived brain-derived neurotrophic factor. Neuron, 18, 257–267.CrossRefGoogle Scholar
Cechetto, D. F. (1995). Supraspinal mechanisms of visceral pain. In Visceral Pain, ed. Gebhardt, G. F.. Seattle: IASP Press, pp. 261–290.Google Scholar
Cechetto, D. F. and Saper, C. B. (1990). Role of the cerebral cortex in autonomic function. In Central Regulation of Autonomic Functions, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 208–223.Google Scholar
Celander, O. (1954). The range of control exercised by the sympatho-adrenal system. Acta Physiol. Scand. Suppl., 116, 1–132.Google Scholar
Cervero, F. (1982). Afferent activity evoked by natural stimulation of the biliary system in the ferret. Pain, 13, 137–151.CrossRefGoogle ScholarPubMed
Cervero, F. (1994). Sensory innervation of the viscera: peripheral basis of visceral pain. Physiol. Rev., 74, 95–138.CrossRefGoogle ScholarPubMed
Cervero, F. (1995). Mechanisms of visceral pain: past and present. In Visceral Pain, ed. Gebhart, G. F.. Seattle: IASP Press, pp. 25–41.Google Scholar
Cervero, F. (1996). Visceral nociceptors. In Neurobiology of Nociceptors, ed. Belmonte, C. and Cervero, F.. Oxford, New York, Toronto: Oxford University Press, pp. 220–240.CrossRefGoogle Scholar
Cervero, F. and Connell, L. A. (1984). Distribution of somatic and visceral primary afferent fibers within the thoracic spinal cord of the cat. J. Comp. Neurol., 230, 88–98.CrossRefGoogle ScholarPubMed
Cervero, F. and Jänig, W. (1992). Visceral nociceptors: a new world order? Trends Neurosci., 15, 374–378.CrossRefGoogle ScholarPubMed
Cervero, F. and Morrison, J. F. B. (eds.) (1986). Visceral Sensation. Prog. Brain Res. Vol. 67, Amsterdam: Elsevier.Google Scholar
Cervero, F. and Sann, H. (1989). Mechanically evoked responses of afferent fibers innervating the guinea-pig's ureter: an in vitro study. J. Physiol., 412, 245–266.CrossRefGoogle Scholar
Cervero, F. and Tattersall, J. E. (1986). Somatic and visceral sensory integration in the thoracic spinal cord. Prog. Brain Res., 67, 189–205.CrossRefGoogle ScholarPubMed
Chan, R. K. and Sawchenko, P. E. (1998). Organization and transmitter specificity of medullary neurons activated by sustained hypertension: implications for understanding baroreceptor reflex circuitry. J. Neurosci., 18, 371–387.CrossRefGoogle ScholarPubMed
Chandler, M. J., Zhang, J., Qin, C. and Foreman, R. D. (2002). Spinal inhibitory effects of cardiopulmonary afferent inputs in monkeys: neuronal processing in high cervical segments. J. Neurophysiol., 87, 1290–1302.CrossRefGoogle ScholarPubMed
Chang, H. S., Staras, K. and Gilbey, M. P. (2000). Multiple oscillators provide metastability in rhythm generation. J. Neurosci., 20, 5135–5143.CrossRefGoogle ScholarPubMed
Chang, H. S., Staras, K., Smith, J. E. and Gilbey, M. P. (1999). Sympathetic neuronal oscillators are capable of dynamic synchronization. J. Neurosci., 19, 3183–3197.CrossRefGoogle ScholarPubMed
Chapleau, M. W. and Abboud, F. (eds.) (2001). Neuro-Cardiovascular Regulation: from Molecules to Man. New York: The New York Academy of Sciences.Google Scholar
Chau, D. and Schramm, L. P. (1997). Sympathetically correlated activity of dorsal horn neurons in spinally transected rats. J. Neurophysiol., 77, 2966–2974.CrossRefGoogle ScholarPubMed
Chau, D., Johns, D. G. and Schramm, L. P. (2000). Ongoing and stimulus-evoked activity of sympathetically correlated neurons in the intermediate zone and dorsal horn of acutely spinalized rats. J. Neurophysiol., 83, 2699–2707.CrossRefGoogle ScholarPubMed
Cheng, Z. and Powley, T. L. (2000). Nucleus ambiguus projections to cardiac ganglia of rat atria: an anterograde tracing study. J. Comp. Neurol., 424, 588–606.3.0.CO;2-7>CrossRefGoogle Scholar
Cheng, Z., Powley, T. L., Schwaber, J. S. and Doyle, F. J., III. (1999). Projections of the dorsal motor nucleus of the vagus to cardiac ganglia of rat atria: an anterograde tracing study. J. Comp. Neurol., 410, 320–341.3.0.CO;2-5>CrossRefGoogle Scholar
Chizh, B. A., Headley, P. M. and Paton, J. F. (1998). Coupling of sympathetic and somatic motor outflows from the spinal cord in a perfused preparation of adult mouse in vitro. J. Physiol., 508, 907–918.CrossRefGoogle Scholar
Choate, J. K., Edwards, F. R., Hirst, G. D. and O'Shea, J. E. (1993a). Effects of sympathetic nerve stimulation on the sino-atrial node of the guinea-pig. J. Physiol., 471, 707–727.CrossRefGoogle Scholar
Choate, J. K., Klemm, M. and Hirst, G. D. S. (1993b). Sympathetic and parasympathetic neuromuscular junctions in the guinea-pig sino-atrial node. J. Auton. Nerv. Syst., 44, 1–15.CrossRefGoogle Scholar
Christensen, J. (1994). The motility of the colon. In Physiology of the Gastrointestinal Tract, ed. Johnson, L. R.. New York: Raven Press, pp. 991–1024.Google Scholar
Christian, E. P. and Weinreich, D. (1988). Long-duration spike afterhyperpolarizations in neurons from the guinea pig superior cervical ganglion. Neurosci. Lett., 84, 191–196.CrossRefGoogle ScholarPubMed
Chrousos, G. P. (1998). Stressors, stress, and neuroendocrine integration of the adaptive response. The 1997 Hans Selye Memorial Lecture. Ann. New York Acad. Sci., 851, 311–335.CrossRefGoogle Scholar
Ciriello, J., Hochstenbach, S. L. and Roder, S. (1994). Central projections of baroreceptor and chemoreceptor afferent fibers in the rat. In Nucleus of the Solitary Tract, ed. Barraco, I. R. A.. Boca Raton: CRC Press, pp. 35–50.Google Scholar
Clayton, E. C. and Williams, C. L. (2000). Adrenergic activation of the nucleus tractus solitarius potentiates amygdala norepinephrine release and enhances retention performance in emotionally arousing and spatial memory tasks. Behav. Brain Res., 112, 151–158.CrossRefGoogle ScholarPubMed
Clement, C. I., Keay, K. A., Owler, B. K. and Bandler, R. (1996). Common patterns of increased and decreased fos expression in midbrain and pons evoked by noxious deep somatic and noxious visceral manipulations in the rat. J. Comp. Neurol., 366, 495–515.3.0.CO;2-#>CrossRefGoogle ScholarPubMed
Clement, C. I., Keay, K. A., Podzebenko, K., Gordon, B. D. and Bandler, R. (2000). Spinal sources of noxious visceral and noxious deep somatic afferent drive onto the ventrolateral periaqueductal gray of the rat. J. Comp. Neurol., 425, 323–344.3.0.CO;2-Z>CrossRefGoogle ScholarPubMed
Clutter, W. E., Bier, D. M., Shah, S. D. and Cryer, P. E. (1980). Epinephrine plasma metabolic clearance rates and physiologic thresholds for metabolic and hemodynamic actions in man. J. Clin. Invest., 66, 94–101.CrossRefGoogle ScholarPubMed
Cochrane, K. L. and Nathan, M. A. (1989). Normotension in conscious rats after placement of bilateral electrolytic lesions in the rostral ventrolateral medulla. J Auton. Nerv. Syst., 26, 199–211.CrossRefGoogle ScholarPubMed
Cochrane, K. L. and Nathan, M. A. (1993). Cardiovascular effects of lesions of the rostral ventrolateral medulla and the nucleus reticularis parvocellularis in rats. J. Auton. Nerv. Syst., 43, 69–81.CrossRefGoogle ScholarPubMed
Coderre, T. J., Basbaum, A. I. and Levine, J. D. (1989). Neural control of vascular permeability: interaction between primary afferents, mast cells, and sympathetic efferents. J. Neurophysiol., 62, 48–58.CrossRefGoogle ScholarPubMed
Coleridge, H. M. and Coleridge, J. C. (1980). Cardiovascular afferents involved in regulation of peripheral vessels. Annu. Rev. Physiol., 42, 413–427.CrossRefGoogle ScholarPubMed
Coleridge, H. M., Coleridge, J. C. G. (1997). Afferent nerves in the airways. In Autonomic Control of the Respiratory System, ed. Barnes, P. J.. Amsterdam: Harwood Academic Publishers GmbH, pp. 39–58.Google Scholar
Coleridge, H. M., Coleridge, J. C., Kaufman, M. P. and Dangel, A. (1981). Operational sensitivity and acute resetting of aortic baroreceptors in dogs. Circ. Res., 48, 676–684.CrossRefGoogle ScholarPubMed
Coleridge, J. C. and Coleridge, H. M. (1984). Afferent vagal C fibre innervation of the lungs and airways and its functional significance. Rev. Physiol. Biochem. Pharmacol., 99, 1–110.CrossRefGoogle ScholarPubMed
Contreras, R. J., Gomez, M. M. and Norgren, R. (1980). Central origins of cranial nerve parasympathetic neurons in the rat. J. Comp. Neurol., 190, 373–394.CrossRefGoogle ScholarPubMed
Cooke, H. J. (1994). Neuroimmune signaling in regulation of intestinal ion transport. Am. J. Physiol., 266, G167–G178.Google ScholarPubMed
Cooke, H. J. (1998). “Enteric tears”: chloride secretion and its neural regulation. News Physiol. Sci., 13, 269–274.Google ScholarPubMed
Cooke, H. J. and Reddix, R. A. (1994). Neural regulation of intestinal electrolyte transport. In Physiology of the Gastrointestinal Tract, 3rd edn., ed. Johnson, L. R.. New York: Raven Press, pp. 2083–2132.Google Scholar
Coolen, L. M., Allard, J., Truit, W. A. and McKenna, K. E. (2004). Central regulation of ejaculation. Physiol. Behav., 83, 203–213.CrossRefGoogle ScholarPubMed
Coonan, E. M., Downie, J. W. and Du, H. J. (1999). Sacral spinal cord neurons responsive to bladder pelvic and perineal inputs in cats. Neurosci. Lett., 260, 137–140.CrossRefGoogle ScholarPubMed
Coote, J. H. (1984). Spinal and supraspinal reflex pathways of cardio-cardiac sympathetic reflexes. Neurosci. Lett., 46, 243–247.CrossRefGoogle ScholarPubMed
Coote, J. H. (1988). The organisation of cardiovascular neurons in the spinal cord. Rev. Physiol. Biochem. Pharmacol., 110, 147–285.CrossRefGoogle ScholarPubMed
Coote, J. H. and Downman, C. B. (1966). Central pathways of some autonomic reflex discharges. J. Physiol., 183, 714–729.CrossRefGoogle ScholarPubMed
Coote, J. H. and Sato, A. (1978). Supraspinal regulation of spinal reflex discharge into cardiac sympathetic nerves. Brain Res., 142, 425–437.CrossRefGoogle ScholarPubMed
Coote, J. H., Macleod, V. H., Fleetwood-Walker, S. M. and Gilbey, M. P. (1981). Baroreceptor inhibition of sympathetic activity at a spinal site. Brain Res., 220, 81–93.CrossRefGoogle Scholar
Costa, M., Brookes, S. J., Steele, P. A.et al. (1996). Neurochemical classification of myenteric neurons in the guinea-pig ileum. Neuroscience, 75, 949–967.CrossRefGoogle ScholarPubMed
Coupland, R. E. (1965). Electron microscopic observations on the structure of the rat adrenal medulla. I. The ultrastructure and organization of chromaffin cells in the normal adrenal medulla. J. Anat., 99, 231–254.Google ScholarPubMed
Cousins, H. M., Edwards, F. R., Hirst, G. D. and Wendt, I. R. (1993). Cholinergic neuromuscular transmission in the longitudinal muscle of the guinea-pig ileum. J. Physiol., 471, 61–86.CrossRefGoogle ScholarPubMed
Cousins, H. M., Edwards, F. R. and Hirst, G. D. (1995). Neuronally released and applied acetylcholine on the longitudinal muscle of the guinea-pig ileum. Neuroscience, 65, 193–207.CrossRefGoogle ScholarPubMed
Coutinho, S. V., Su, X., Sengupta, J. N. and Gebhart, G. F. (2000). Role of sensitized pelvic nerve afferents from the inflamed rat colon in the maintenance of visceral hyperalgesia. Prog. Brain Res., 129, 375–387.CrossRefGoogle Scholar
Cowley, A. W. Jr. (1992). Long-term control of arterial blood pressure. Physiol. Rev., 72, 231–300.CrossRefGoogle ScholarPubMed
Cowley, A. W. Jr., Liard, J. F. and Guyton, A. C. (1973). Role of the baroreceptor reflex in daily control of arterial blood pressure and other variables in dogs. Circ. Res., 32, 564–576.CrossRefGoogle ScholarPubMed
Cox, G. E., Jordan, D., Paton, J. F., Spyer, K. M. and Wood, L. M. (1987). Cardiovascular and phrenic nerve responses to stimulation of the amygdala central nucleus in the anaesthetized rabbit. J. Physiol., 389, 541–556.CrossRefGoogle ScholarPubMed
Craig, A. D. (1996). An ascending general homeostatic afferent pathway originating in lamina I. Prog. Brain Res., 107, 225–243.CrossRefGoogle ScholarPubMed
Craig, A. D. (2002). How do you feel? Interoception: the sense of the physiological condition of the body. Nat. Rev. Neurosci., 3, 655–666.CrossRefGoogle Scholar
Craig, A. D. (2003a). Pain mechanisms: labeled lines versus convergence in central processing. Annu. Rev. Neurosci., 26, 1–30.CrossRefGoogle Scholar
Craig, A. D. (2003b). A new view of pain as a homeostatic emotion. Trends Neurosci., 26, 303–307.CrossRefGoogle Scholar
Craig, A. D. (2003c). Interoception: the sense of the physiological condition of the body. Curr. Opin. Neurobiol., 13, 500–505.CrossRefGoogle Scholar
Craig, A. D. (2004a). Lamina I, but not lamina V, spinothalamic neurons exhibit responses that correspond with burning pain. J. Neurophysiol., 92, 2604–2609.CrossRefGoogle Scholar
Craig, A. D. (2004b). Distribution of trigeminothalamic and spinothalamic lamina I terminations in the macaque monkey. J. Comp. Neurol., 477, 119–148.CrossRefGoogle Scholar
Craig, A. D. and Blomqvist, A. (2002). Is there a specific lamina I spinothalamocortical pathway for pain and temperature sensations in primates?J. Pain, 3, 95–101.CrossRefGoogle Scholar
Craig, A. D. and Kniffki, K. D. (1985). Spinothalamic lumbosacral lamina I cells responsive to skin and muscle stimulation in the cat. J. Physiol., 365, 197–221.CrossRefGoogle ScholarPubMed
Craig, A. D. and Mense, S. (1983). The distribution of afferent fibers from the gastrocnemius-soleus muscle in the dorsal horn of the cat, as revealed by the transport of horseradish peroxidase. Neurosci. Lett., 41, 233–238.CrossRefGoogle ScholarPubMed
Craig, A. D., Heppelmann, B. and Schaible, H. G. (1988). The projection of the medial and posterior articular nerves of the cat's knee to the spinal cord. J. Comp. Neurol., 276, 279–288.CrossRefGoogle ScholarPubMed
Craig, A. D., Bushnell, M. C., Zhang, E. T. and Blomqvist, A. (1994). A thalamic nucleus specific for pain and temperature sensation. Nature, 372, 770–773.CrossRefGoogle ScholarPubMed
Craig, A. D., Krout, K. and Andrew, D. (2001). Quantitative response characteristics of thermoreceptive and nociceptive lamina I spinothalamic neurons in the cat. J. Neurophysiol., 86, 1459–1480.CrossRefGoogle ScholarPubMed
Cravo, S. L., Morrison, S. F. and Reis, D. J. (1991). Differentiation of two cardiovascular regions within caudal ventrolateral medulla. Am. J. Physiol., 261, R985–R994.Google ScholarPubMed
Crawford, J. P. and Frankel, H. L. (1971). Abdominal ‘visceral’ sensation in human tetraplegia. Paraplegia, 9, 153–158.Google ScholarPubMed
Critchley, H. and Dolan, R. J. (2004). Central representation of autonomic states. In Human Brain Function. 2nd edn., eds. Frackowiak, R. S. J., Friston, K. J., Frith, C. D.et al. Amsterdam: Elsevier Academic Press, pp. 397–417.Google Scholar
Crowcroft, P. J., Holman, M. E. and Szurszewski, J. H. (1971). Excitatory input from the distal colon to the inferior mesenteric ganglion in the guinea-pig. J. Physiol., 219, 443–461.CrossRefGoogle ScholarPubMed
Cryer, P. E. (1980). Physiology and pathophysiology of the human sympathoadrenal neuroendocrine system. New Engl. J. Med., 303, 436–444.Google ScholarPubMed
Cunnane, T. C. and Stjärne, L. (1982). Secretion of transmitter from individual varicosities of guinea-pig and mouse vas deferens: all-or-none and extremely intermittent. Neuroscience, 7, 2565–2576.CrossRefGoogle ScholarPubMed
Cyon, E. and Ludwig, C. (1866). Die Reflexe eines der sensiblen Nerven des Herzens auf die motorischen der Blutgefässe [The reflex of one of the heart nerves on the motor nerves to blood vessels]. Ber. Sächs. Ges. Wiss., 18, 307–328.Google Scholar
Czachurski, J., Dembowsky, K., Seller, H., Nobiling, R. and Taugner, R. (1988). Morphology of electrophysiologically identified baroreceptor afferents and second order neurones in the brainstem of the cat. Arch. Ital. Biol., 126, 129–144.Google ScholarPubMed
Dado, R. J., Katter, J. T. and Giesler, G. J. Jr. (1994). Spinothalamic and spinohypothalamic tract neurons in the cervical enlargement of rats. I. Locations of antidromically identified axons in the thalamus and hypothalamus. J. Neurophysiol., 71, 959–980.CrossRefGoogle Scholar
Dail, W. G. (1993). Autonomic innervation of male reproductive genitalia. In The Autonomic Nervous System, Vol. 3, Nervous Control of the Urogenital System, ed. Maggi, C. A.. Chur, Switzerland: Harwood Academic Publishers, pp. 69–101.Google Scholar
Dalsgaard, C. J. and Elfvin, L. G. (1982). Structural studies on the connectivity of the inferior mesenteric ganglion of the guinea pig. J. Auton. Nerv. Syst., 5, 265–278.CrossRefGoogle ScholarPubMed
Dalsgaard, C. J., Hökfelt, T., Elfvin, L. G., Skirboll, L. and Emson, P. (1982). Substance P-containing primary sensory neurons projecting to the inferior mesenteric ganglion: evidence from combined retrograde tracing and immunohistochemistry. Neuroscience, 7, 647–654.CrossRefGoogle ScholarPubMed
Daly, M.d. (1991). Some reflex cardioinhibitory responses in the cat and their modulation by central inspiratory neuronal activity. J. Physiol., 439, 559–577.CrossRefGoogle ScholarPubMed
Daly, M.d. and Robinson, B. H. (1968). An analysis of the reflex systemic vasodilator response elicited by lung inflation in the dog. J. Physiol., 195, 387–406.CrossRefGoogle ScholarPubMed
Daly, M.d., Hazzledine, J. L. and Ungar, A. (1967). The reflex effects of alterations in lung volume on systemic vascular resistance in the dog. J. Physiol., 188, 331–351.CrossRefGoogle Scholar
Daly, M.d., Ward, J. and Wood, L. M. (1987). The peripheral chemoreceptors and cardiovascular-respiratory integration. In The Neurobiology of the Cardiorespiratory System, ed. Taylor, E. W.. Manchester: Manchester University Press, pp. 342–368.Google Scholar
Damasio, A. R. (1994). Descartes' Error: Emotion, Reason and the Human Brain. New York: Avon Books.Google Scholar
Damasio, A. R. (1999). The Feeling of What Happens: Body and Emotion in the Making of Consciousness. Harcourt Brace: New York.Google Scholar
Damasio, A. R., Adolphs, R. and Damasio, H. (2003). The contributions of the lesion method to the functional neuroanatomy of emotion. In Handbook of Affective Sciences, eds. Davidson, R. J., Scherer, K. R. and Goldsmith, H. H.. Oxford, New York: Oxford University Press, pp. 66–92.Google Scholar
Dampney, R. A. (1981). Brain stem mechanisms in the control of arterial pressure. Clin. Exp. Hypertens., 3, 379–391.CrossRefGoogle ScholarPubMed
Dampney, R. A. (1994). Functional organization of central pathways regulating the cardiovascular system. Physiol. Rev., 74, 323–364.CrossRefGoogle ScholarPubMed
Dampney, R. A. and Horiuchi, J. (2003). Functional organisation of central cardiovascular pathways: studies using c-fos gene expression. Prog. Neurobiol., 71, 359–384.CrossRefGoogle ScholarPubMed
Dampney, R. A. and McAllen, R. M. (1988). Differential control of sympathetic fibres supplying hindlimb skin and muscle by subretrofacial neurones in the cat. J. Physiol., 395, 41–56.CrossRefGoogle ScholarPubMed
Dampney, R. A. and Moon, E. A. (1980). Role of ventrolateral medulla in vasomotor response to cerebral ischemia. Am. J. Physiol., 239, H349–H358.Google ScholarPubMed
Dampney, R. A., Goodchild, A. K., Robertson, L. G. and Montgomery, W. (1982). Role of ventrolateral medulla in vasomotor regulation: a correlative anatomical and physiological study. Brain Res., 249, 223–235.CrossRefGoogle ScholarPubMed
Dampney, R. A., Goodchild, A. K. and Tan, E. (1985). Vasopressor neurons in the rostral ventrolateral medulla of the rabbit. J. Auton. Nerv. Syst., 14, 239–254.CrossRefGoogle ScholarPubMed
Dampney, R. A., Tagawa, T., Horiuchi, J.et al. (2000). What drives the tonic activity of presympathetic neurons in the rostral ventrolateral medulla?Clin. Exp. Pharmacol. Physiol., 27, 1049–1053.CrossRefGoogle ScholarPubMed
Dampney, R. A., Horiuchi, J., Tagawa, T.et al. (2003). Medullary and supramedullary mechanisms regulating sympathetic vasomotor tone. Acta Physiol. Scand., 177, 209–218.CrossRefGoogle ScholarPubMed
Dantzer, R., Bluthe, R. M., Gheusi, G.et al. (1998). Molecular basis of sickness behavior. Ann. New York Acad. Sci., 856, 132–138.CrossRefGoogle ScholarPubMed
Dantzer, R., Konsman, J. P., Bluthe, R. M. and Kelley, K. W. (2000). Neural and humoral pathways of communication from the immune system to the brain: parallel or convergent?Auton. Neurosci., 85, 60–65.CrossRefGoogle ScholarPubMed
Darwin, C. (1998) [1872]. The Expression of the Emotions in Man and Animals. With an Introduction, Afterword and Commentaries by Paul Ekman, 3rd edn. London: Harper Collins.Google Scholar
Davidson, R. J., Scherer, K. R. and Goldsmith, H. H. (eds.) (2003). Handbook of Affective Sciences. New York, Oxford: Oxford University Press.Google Scholar
Davies, P. J., Ireland, D. R. and McLachlan, E. M. (1996). Sources of Ca2+ for different Ca2 + -activated K+ conductances in neurones of the rat superior cervical ganglion. J. Physiol., 495, 353–366.CrossRefGoogle ScholarPubMed
Davies, P. J., Ireland, D. R., Martinez-Pinna, J. and McLachlan, E. M. (1999). Electrophysiological roles of L-type channels in different classes of guinea pig sympathetic neuron. J. Neurophysiol., 82, 818–828.CrossRefGoogle ScholarPubMed
Davis, K. D., Meyer, R. A. and Campbell, J. N. (1993). Chemosensitivity and sensitization of nociceptive afferents that innervate the hairy skin of monkey. J. Neurophysiol., 69, 1071–1081.CrossRefGoogle ScholarPubMed
Davis, M. J. and Hill, M. A. (1999). Signaling mechanisms underlying the vascular myogenic response. Physiol. Rev., 79, 387–423.CrossRefGoogle ScholarPubMed
Groat, W. C. (1976). Mechanisms underlying recurrent inhibition in the sacral parasympathetic outflow to the urinary bladder. J. Physiol., 257, 503–513.CrossRefGoogle ScholarPubMed
Groat, W. C. (1987). Neuropeptides in pelvic afferent pathways. Experientia, 43, 801–813.CrossRefGoogle ScholarPubMed
de Groat, W. C. (1989). Neuropeptides in pelvic afferent pathways. In Regulatory Peptides, ed. Polak, J. M.. Basel: Birkhauser Verlag AG, pp. 334–361.CrossRefGoogle Scholar
de Groat, W. C. (2002). Neural control of the urinary bladder and sexual organs. In Autonomic Failure, 4th edn., eds. Mathias, C. J. and Bannister, R.. New York, Oxford: Oxford University Press, pp. 151–165.Google ScholarPubMed
de Groat, W. C. and Booth, A. M. (1993). Neural control of penile erection. In The Autonomic Nervous System, Vol. 3, Nervous Control of the Urogenital System, ed. Maggi, C. A.. Chur, Switzerland: Harwood Academic Publishers, pp. 467–524.Google Scholar
Groat, W. C. and Ryall, R. W. (1968). Recurrent inhibition in sacral parasympathetic pathways to the bladder. J. Physiol., 196, 579–591.CrossRefGoogle Scholar
Groat, W. C., Booth, A. M., Milne, R. J. and Roppolo, J. R. (1982). Parasympathetic preganglionic neurons in the sacral spinal cord. J. Auton. Nerv. Syst., 5, 23–43.CrossRefGoogle ScholarPubMed
de Groat, W. C., Booth, A. M. and Yoshimura, N. (1993). Neurophysiology of micturition and its modification in animal models in human disease. In The Autonomic Nervous System, Vol. 3, Nervous Control of the Urogenital System. ed. Maggi, C. A.. Chur, Switzerland: Harwood Academic Publishers, pp. 227–290.Google Scholar
Groat, W. C., Vizzard, M. A., Araki, I. and Roppolo, J. H. (1996). Spinal interneurons and preganglionic neurons in sacral autonomic reflex pathways. Prog. Brain Res., 107, 97–111.CrossRefGoogle Scholar
Groat, W. C., Araki, I., Vizzard, M. A.et al. (1998). Developmental and injury induced plasticity in the micturition reflex pathway. Behav. Brain Res., 92, 127–140.CrossRefGoogle ScholarPubMed
Groat, W. C., Fraser, M. O., Yoshiyama, M.et al. (2001). Neural control of the urethra. Scand. J. Urol. Nephrol., Suppl (207), 35–43.Google ScholarPubMed
Delius, W., Hagbarth, K. E., Hongell, A. and Wallin, B. G. (1972). General characteristics of sympathetic activity in human muscle nerves. Acta Physiol. Scand., 84, 65–81.CrossRefGoogle ScholarPubMed
Dembowsky, K., Czachurski, J. and Seller, H. (1985). Morphology of sympathetic preganglionic neurons in the thoracic spinal cord of the cat: an intracellular horseradish peroxidase study. J. Comp. Neurol., 238, 453–465.CrossRefGoogle Scholar
Demir, S. S., Clark, J. W. and Giles, W. R. (1999). Parasympathetic modulation of sinoatrial node pacemaker activity in rabbit heart: a unifying model. Am. J. Physiol., 276, H2221–H2244.Google ScholarPubMed
Denton, K. M., Luff, S. E., Shweta, A. and Anderson, W. P. (2004). Differential neural control of glomerular ultrafiltration. Clin. Exp. Pharmacol. Physiol., 31, 380–386.CrossRefGoogle ScholarPubMed
Deuchars, S. A., Spyer, K. M. and Gilbey, M. P. (1997). Stimulation within the rostral ventrolateral medulla can evoke monosynaptic γ-aminobutyric acidergic inhibitory postsynaptic potentials in sympathetic preganglionic neurons in vitro. J. Neurophysiol., 77, 229–235.CrossRefGoogle Scholar
Deuchars, J., Li, Y. W., Kasparov, S. and Paton, J. F. (2000). Morphological and electrophysiological properties of neurones in the dorsal vagal complex of the rat activated by arterial baroreceptors. J. Comp. Neurol., 417, 233–249.3.0.CO;2-V>CrossRefGoogle ScholarPubMed
Deuchars, S. A., Brooke, R. E. and Deuchars, J. (2001a). Adenosine A1 receptors reduce release from excitatory but not inhibitory synaptic inputs onto lateral neurons. J. Neurosci., 21, 6308–6320.CrossRefGoogle Scholar
Deuchars, S. A., Brooke, R. E., Frater, B. and Deuchars, J. (2001b). Properties of interneurones in the intermediolateral cell column of the rat spinal cord: role of the potassium channel subunit Kv3.1. Neuroscience, 106, 433–446.CrossRefGoogle Scholar
Deuchars, S. A., Milligan, C. J., Stornetta, R. L. and Deuchars, J. (2005). γ-aminobutyric acidergic neurons in the central region of the spinal cord: a novel substrate for sympathetic inhibition. J. Neurosci., 25, 1063–1070.CrossRefGoogle Scholar
DiBona, G. F. and Kopp, U. C. (1997). Neural control of renal function. Physiol. Rev., 77, 75–197.CrossRefGoogle ScholarPubMed
Dickens, E. J., Hirst, G. D. and Tomita, T. (1999). Identification of rhythmically active cells in guinea-pig stomach. J. Physiol., 514, 515–531.CrossRefGoogle ScholarPubMed
Dietz, N. M., Rivera, J. M., Eggener, S. E.et al. (1994). Nitric oxide contributes to the rise in forearm blood flow during mental stress in humans. J. Physiol., 480, 361–368.CrossRefGoogle ScholarPubMed
Dittmar, C. (1873). Über die Lage des sogenannten Gefässcentrums in der Medulla oblongata [On the location of the so-called vascular center in the medulla oblongata]. Sitzungsber. Akad. Wiss. Wien, Math.-Naturw., Abt. 2, 25, 449–469.Google Scholar
Dobbins, E. G. and Feldman, J. L. (1994). Brainstem network controlling descending drive to phrenic motoneurons in rat. J. Comp. Neurol., 347, 64–86.CrossRefGoogle ScholarPubMed
Dockray, G. J., Green, T. and Varro, A. (1989). The afferent peptidergic innervation of the upper gastrointestinal tract. In Nerves and the Gastrointestinal Tract, eds. Singer, M. V. and Goebel, H.. Dordrecht: Kluwer Academic Publishers, pp. 105–122.Google Scholar
Dodt, C., Gunnarsson, T., Elam, M., Karlsson, T. and Wallin, B. G. (1995). Central blood volume influences sympathetic sudomotor nerve traffic in warm humans. Acta Physiol. Scand., 155, 41–51.CrossRefGoogle ScholarPubMed
Dodt, C., Lönnroth, P., Fehm, H. L. and Elam, M. (1999). Intraneural stimulation elicits an increase in subcutaneous interstitial glycerol levels in humans. J. Physiol., 521, 545–552.CrossRefGoogle ScholarPubMed
Donnerer, J., Schuligoi, R. and Stein, C. (1992). Increased content and transport of substance P and calcitonin gene-related peptide in sensory nerves innervating inflamed tissue: evidence for a regulatory function of nerve growth factor in vivo. Neuroscience, 49, 693–698.CrossRefGoogle ScholarPubMed
Dorward, P. K., Andresen, M. C., Burke, S. L., Oliver, J. R. and Korner, P. I. (1982). Rapid resetting of the aortic baroreceptors in the rabbit and its implications for short-term and longer term reflex control. Circ. Res., 50, 428–439.CrossRefGoogle ScholarPubMed
Dorward, P. K., Burke, S. L., Jänig, W. and Cassell, J. (1987). Reflex responses to baroreceptor, chemoreceptor and nociceptor inputs in single renal sympathetic neurones in the rabbit and the effects of anaesthesia on them. J. Auton. Nerv. Syst., 18, 39–54.CrossRefGoogle Scholar
Dostrovsky, J. O. and Craig, A. D. (2005). Ascending projection systems. In Wall and Melzack's Textbook of Pain. 5th edn., ed. McMahon, S. B. and Koltzenburg, M.. Edinburgh: Elsevier Churchill Livingstone, pp. 187–204.Google Scholar
Downing, J. E. and Miyan, J. A. (2000). Neural immunoregulation: emerging roles for nerves in immune homeostasis and disease. Immunol. Today, 21, 281–289.CrossRefGoogle ScholarPubMed
Drummond, P. D. (1995). Mechanisms of physiological gustatory sweating and flushing in the face. J. Auton. Nerv. Syst., 52, 117–124.CrossRefGoogle Scholar
Duan, Y. F., Winters, R., McCabe, P. M.et al. (1996). Behavioral characteristics of defense and vigilance reactions elicited by electrical stimulation of the hypothalamus in rabbits. Behav. Brain Res., 81, 33–41.CrossRefGoogle ScholarPubMed
Duan, Y. F., Winters, R., McCabe, P. M.et al. (1997). Functional relationship between the hypothalamic vigilance area and periaqueductal grey vigilance area. Physiol. Behav., 62, 675–679.CrossRefGoogle ScholarPubMed
Dun, N. J. (1983). Peptide hormones and transmission in sympathetic ganglia. In Autonomic Ganglia, ed. Elfvin, L. G.. Chichester: J. Wiley & Sons, pp. 345–366.Google Scholar
Dunn, W. R., Brock, J. A. and Hardy, T. A. (1999). Electrochemical and electrophysiological characterization of neurotransmitter release from sympathetic nerves supplying rat mesenteric arteries. Br. J. Pharmacol., 128, 174–180.CrossRefGoogle ScholarPubMed
Dworkin, B. R. (1993). Learning and Physiological Regulation. Chicago: The University of Chicago Press.Google Scholar
Dworkin, B. R. (2000). Interoception. In Handbook of Psychophysiology, 2nd edn., ed. Cacioppo, J. T., Tassinary, L. G. and Berntson, G. G.. Cambridge: Cambridge University Press, pp. 482–506.Google Scholar
Ebbeson, S. O. E. (1968a). Quantitative studies of superior cervical sympathetic ganglia in a variety of primates including man. I. The ratio of preganglionic neurons. J. Morphol., 124, 117–132.CrossRefGoogle Scholar
Ebbeson, S. O. E. (1968b). Quantitative studies of superior cervical sympathetic ganglia in a variety of primates including man. II. Neuronal packing density. J. Morphol., 124, 181–186.CrossRefGoogle Scholar
Eckberg, D. L. (2003). The human respiratory gate. J. Physiol., 548, 339–352.Google ScholarPubMed
Eckberg, D. L. and Sleight, P. (1992). Human Baroreceptor Reflexes in Health and Disease. Oxford: Clarendon Press.Google Scholar
Eckberg, D. L., Nerhed, C. and Wallin, B. G. (1985). Respiratory modulation of muscle sympathetic and vagal cardiac outflow in man. J. Physiol., 365, 181–196.CrossRefGoogle ScholarPubMed
Eckberg, D. L., Rea, R. F., Andersson, O. K.et al. (1988). Baroreflex modulation of sympathetic activity and sympathetic neurotransmitters in humans. Acta Physiol. Scand., 133, 221–231.CrossRefGoogle ScholarPubMed
Ectors, L. (1941). Contribution á l'étude des réactions pilomotrices. Arch. Int. Physiol., 51, 443–455.Google Scholar
Edwards, F. R., Bramich, N. J. and Hirst, G. D. (1993). Analysis of the effects of vagal stimulation on the sinus venous of the toad. Philos. Trans. R. Soc. London B Biol. Sci., 341, 149–162.CrossRefGoogle ScholarPubMed
Edwards, F. R., Hirst, G. D., Klemm, M. F. and Steele, P. A. (1995). Different types of ganglion cell in the cardiac plexus of guinea-pigs. J. Physiol., 486, 453–471.CrossRefGoogle ScholarPubMed
Edwards, S. L., Anderson, C. R., Southwell, B. R. and McAllen, R. M. (1996). Distinct preganglionic neurons innervate noradrenaline and adrenaline cells in the cat adrenal medulla. Neuroscience, 70, 825–832.CrossRefGoogle ScholarPubMed
Ekman, P. (1992). Facial expression of emotion: new findings, new questions. Psychol. Sci., 3, 34–38.CrossRefGoogle Scholar
Ekman, P. and Davidson, R. J. (eds.) (1994). The Nature of Emotions. Oxford: Oxford University Press.Google Scholar
Ekman, P., Levenson, R. W. and Friesen, M. V. (1983). Autonomic nervous system activity distinguishes between emotions. Science, 221, 1208–1210.CrossRefGoogle ScholarPubMed
Elfvin, L. G. (ed.) (1983). Autonomic Ganglia. Chichester: J. Wiley & Sons.Google Scholar
Elfvin, L. G., Lindh, B. and Hökfelt, T. (1993). The chemical neuroanatomy of sympathetic ganglia. Annu. Rev. Neurosci., 16, 471–507.CrossRefGoogle ScholarPubMed
Eliasson, S., Folkow, B., Lindgren, P. and Uvnäs, B. (1951). Activation of sympathetic vasodilator nerves to the skeletal muscles in the cat by hypothalamic stimulation. Acta Physiol. Scand., 23, 333–351.CrossRefGoogle ScholarPubMed
Ellison, G. D. and Zanchetti, A. (1973). Diffuse and specific activation of sympathetic cholinergic fibers of the cat. Am. J. Physiol., 225, 142–149.Google ScholarPubMed
Elsner, R., Franklin, D. L., Citters, R. L. and Kenney, D. W. (1966). Cardiovascular defense against asphyxia. Science, 153, 941–949.CrossRefGoogle ScholarPubMed
Emch, G. S., Hermann, G. E. and Rogers, R. C. (2000). tumor necrosis factor-alpha activates solitary nucleus neurons responsive to gastric distension. Am. J. Physiol. Gastrointest. Liver Physiol., 279, G582–G586.CrossRefGoogle ScholarPubMed
Emch, G. S., Hermann, G. E. and Rogers, R. C. (2002). Tumor necrosis factor-alpha inhibits physiologically identified dorsal motor nucleus neurons in vivo. Brain Res., 951, 311–315.CrossRefGoogle ScholarPubMed
Enquist, L. W. and Card, J. P. (2003). Recent advances in the use of neurotropic viruses for circuit analysis. Curr. Opin. Neurobiol., 13, 603–606.CrossRefGoogle ScholarPubMed
Eppel, G. A., Malpas, S. C., Denton, K. M. and Evans, R. G. (2004). Neural control of renal medullary perfusion. Clin. Exp. Pharmacol. Physiol., 31, 387–396.CrossRefGoogle ScholarPubMed
Ernsberger, U. (2000). Evidence for an evolutionary conserved role of bone morphogenetic protein growth factors and phox2 transcription factors during noradrenergic differentiation of sympathetic neurons. Induction of a putative synexpression group of neurotransmitter-synthesizing enzymes. Eur. J. Biochem., 267, 6976–6981.CrossRefGoogle ScholarPubMed
Ernsberger, U. (2001). The development of postganglionic sympathetic neurons: coordinating neuronal differentiation and diversification. Auton. Neurosci., 94, 1–13.CrossRefGoogle ScholarPubMed
Ernsberger, U., Esposito, L., Partimo, S.et al. (2005). Expression of neuronal markers suggests heterogeneity of chick sympathoadrenal cells prior to invasion of the adrenal anlagen. Cell Tissue Res., 319, 1–13.CrossRefGoogle ScholarPubMed
Esler, M., Jennings, G., Lambert, G.et al. (1990). Overflow of catecholamine neurotransmitters to the circulation: source, fate, and functions. Physiol. Rev., 70, 963–985.CrossRefGoogle ScholarPubMed
Evans, R. J. and Cunnane, T. C. (1992). Relative contributions of adenosine triphosphate and noradrenaline to the nerve evoked contraction of the rabbit jejunal artery. Dependence on stimulation parameters. Naunyn-Schmiedeberg's Arch. Pharmacol., 345, 424–430.CrossRefGoogle ScholarPubMed
Evans, R. J. and Surprenant, A. (1992). Vasoconstriction of guinea-pig submucosal arterioles following sympathetic nerve stimulation is mediated by the release of adenosine triphosphate. Br. J. Pharmacol., 106, 242–249.CrossRefGoogle Scholar
Ezure, K., Tanaka, I. and Kondo, M. (2003). Glycine is used as a transmitter by decrementing expiratory neurons of the ventrolateral medulla in the rat. J. Neurosci., 23, 8941–8948.CrossRefGoogle ScholarPubMed
Fagius, J. and Sundlöf, G. (1986). The diving response in man: effects on sympathetic activity in muscle and skin nerve fascicles. J. Physiol., 377, 429–443.CrossRefGoogle ScholarPubMed
Fagius, J. and Wallin, B. G. (1993). Long-term variability and reproducibility of resting human muscle nerve sympathetic activity at rest, as reassessed after a decade. Clin. Auton. Res., 3, 201–205.CrossRefGoogle Scholar
Fagius, J., Wallin, B. G., Sundlöf, G., Nerhed, C. and Englesson, S. (1985). Sympathetic outflow in man after anaesthesia of the glossopharyngeal and vagus nerves. Brain, 108, 423–438.CrossRefGoogle ScholarPubMed
Farkas, E., Jansen, A. S. and Loewy, A. D. (1998). Periaqueductal gray matter input to cardiac-related sympathetic premotor neurons. Brain Res., 792, 179–192.CrossRefGoogle ScholarPubMed
Feigl, E. O. (1998). Neural control of coronary blood flow. J. Vasc. Res., 35, 85–92.CrossRefGoogle ScholarPubMed
Feldberg, W. and Guertzenstein, P. G. (1976). Vasodepressor effects obtained by drugs acting on the ventral surface of the brain stem. J. Physiol., 258, 337–355.CrossRefGoogle ScholarPubMed
Feldman, J. L. (1986). Neurophysiology of breathing in mammals. In Handbook of Physiology, The Nervous System, Vol. IV, ed. Mountcastle, V. B. and Bloom, F. E., Bethesda: American Physiological Society, pp. 463–524.Google Scholar
Feldman, J. L. and McCrimmon, D. R. (2003). Neural control of breathing. In Fundamental Neuroscience, 2nd edn., ed. Squire, L. R., Bloom, F. E., McConnell, S. K.et al. San Diego: Academic Press, pp. 967–990.Google Scholar
Feldman, J. L., Mitchell, G. S. and Nattie, E. E. (2003). Breathing: rhythmicity, plasticity, chemosensitivity. Annu. Rev. Neurosci., 26, 239–266.CrossRefGoogle Scholar
Fernandez de Molina, A. and Hunsperger, R. W. (1962). Organization of the subcortical system governing defence and flight reactions in the cat. J. Physiol., 160, 200–213.CrossRefGoogle ScholarPubMed
Fielden, R., Sutton, T. J. and Taylor, E. M. (1980). Effect of tilting on the pressor responses to McN-A-343, a muscarinic sympathetic ganglion stimulant. Br. J. Pharmacol., 71, 287–295.CrossRefGoogle ScholarPubMed
Fields, H. L., Basbaum, A. I. and Heinricher, M. M. (2005). Central nervous system mechanisms of pain modulation. In Wall and Melzack's Textbook of Pain, 5th edn., eds. McMahon, S. B. and Koltzenburg, M.. Amsterdam, Edinburgh: Elsevier Churchill Livingstone, pp. 125–142.Google Scholar
Fleming, W. W. and Westfall, D. P. (1988). Adaptive supersensitivity. In Handbook of Experimental Pharmacology, Vol. 90/I, Catecholamines I, ed. Trendelenburg, U. and Weiner, N.. Berlin, Heidelberg, New York: Springer-Verlag, pp. 509–559.Google Scholar
Floyd, N. S., Price, J. L., Ferry, A. T., Keay, K. A. and Bandler, R. (2000). Orbitomedial prefrontal cortical projections to distinct longitudinal columns of the periaqueductal gray in the rat. J. Comp. Neurol., 422, 556–578.3.0.CO;2-U>CrossRefGoogle ScholarPubMed
Floyd, N. S., Price, J. L., Ferry, A. T., Keay, K. A. and Bandler, R. (2001). Orbitomedial prefrontal cortical projections to hypothalamus in the rat. J. Comp. Neurol., 432, 307–328.CrossRefGoogle ScholarPubMed
Foerster, O. (1927). Die Leitungsbahnen des Schmerzgefühls und die chirurgische Behandlung der Schmerzzustände [Pain Pathways and the Surgical Treatment of Pain]. Berlin, Wien: Urban und Schwarzenberg.Google Scholar
Fogel, R., Zhang, X. and Renehan, W. E. (1996). Relationships between the morphology and function of gastric and intestinal distention-sensitive neurons in the dorsal motor nucleus of the vagus. J. Comp. Neurol., 364, 78–91.3.0.CO;2-P>CrossRefGoogle ScholarPubMed
Folkow, B. (1955). The control of blood vessels. Physiol. Rev., 35, 629–663.CrossRefGoogle ScholarPubMed
Folkow, B. (1987). Psychosocial and central nervous influences in primary hypertension. Circulation, 76, I10–I19.Google ScholarPubMed
Folkow, B. (2000). Perspectives on the integrative function of the “sympatho-adrenomedullary system”. Auton. Neurosci., 83, 101–115.CrossRefGoogle Scholar
Folkow, B. and Euler, U. S. (1954). Selective activation of noradrenaline and adrenaline producing cells in the cat's adrenal gland by hypothalamic stimulation. Circ. Res., 2, 191–195.CrossRefGoogle ScholarPubMed
Folkow, B. and Neil, E. (1971). Circulation. New York: Oxford University Press.Google ScholarPubMed
Folkow, B. and Nilsson, H. (1997). Transmitter release at adrenergic nerve endings: Total exocytosis or fractional release?News Physiol. Sci., 12, 32–36.Google Scholar
Folkow, B., Schmidt, T. and Uvnäs-Moberg, K. (eds.) (1997). Stress, health and the social environment. Acta Physiol. Scand., 161, Suppl. 640, 1–179.Google Scholar
Forehand, C. J. (1987). Ultrastructural analysis of the distribution of synaptic boutons from labeled preganglionic axons on rabbit ciliary neurons. J. Neurosci., 7, 3274–3281.CrossRefGoogle ScholarPubMed
Forehand, C. J. and Purves, D. (1984). Regional innervation of rabbit ciliary ganglion cells by the terminals of preganglionic axons. J. Neurosci., 4, 1–12.CrossRefGoogle ScholarPubMed
Foreman, R. D. (1989). Organization of the spinothalamic tract as a relay for cardiopulmonary sympathetic afferent fiber activity. Prog. Sens. Physiol., 9, 1–51.CrossRefGoogle Scholar
Foreman, R. D. (1999). Mechanisms of cardiac pain. Annu. Rev. Physiol., 61, 143–167.CrossRefGoogle ScholarPubMed
Foster, M. (1879). Textbook of Physiology. In The Discovery of Reflexes, ed. Liddell, E. G. T.. Oxford: Clarendon Press, pp. 98–101.Google Scholar
Fox, E. A. and Powley, T. L. (1985). Longitudinal columnar organization within the dorsal motor nucleus represents separate branches of the abdominal vagus. Brain Res., 341, 269–282.CrossRefGoogle ScholarPubMed
Fox, E. A. and Powley, T. L. (1992). Morphology of identified preganglionic neurons in the dorsal motor nucleus of the vagus. J. Comp. Neurol., 322, 79–98.CrossRefGoogle ScholarPubMed
Frayn, K. N. and MacDonald, I. A. (1996). Adipose tissue circulation. In The Autonomic Nervous System, Vol. 8, Nervous Control of Blood Vessels, ed. Bennett, T. and Gardiner, S. M.. Amsterdam: Harwood Academic Publishers, pp. 505–539.Google Scholar
Freeman, R. and Rutkove, S. (2000). Syncope. In Handbook of Clinical Neurology, Vol. 75, The Autonomic Nervous System, part II: Dysfunctions, ed. Appenzeller, O.. Amsterdam: Elsevier, pp. 203–228.Google Scholar
Freyburger, W. A., Gruhzit, C. C. and Moe, G. K. (1950a). Pressor pathways not blocked by tetraethylammonium. Am. J. Physiol., 163, 290–293.Google Scholar
Freyburger, W. A., Gruhzit, C. C., Rennick, B. R. and Moe, G. K. (1950b). Action of tetraethylammonium on pressor response to asphyxia. Am. J. Physiol., 163, 554–560.Google Scholar
Fukai, K. and Fukuda, H. (1985). Three serial neurones in the innervation of the colon by the sacral parasympathetic nerve of the dog. J. Physiol., 362, 69–78.CrossRefGoogle ScholarPubMed
Fukami, H. and Bradley, R. M. (2005). Biophysical and morphological properties of parasympathetic neurons controlling the parotid and von Ebner salivary glands in rats. J. Neurophysiol., 93, 678–686.CrossRefGoogle ScholarPubMed
Fukuda, A., Minami, T., Nabekura, J. and Oomura, Y. (1987). The effects of noradrenaline on neurones in the rat dorsal motor nucleus of the vagus, in vitro. J. Physiol., 393, 213–231.CrossRefGoogle ScholarPubMed
Fulton, J. F. (1949). Physiology of the Nervous System, 3rd edn. New York: Oxford University Press.Google Scholar
Furness, J. B. (2005). The Enteric Nervous System. Oxford: Blackwell Science Ltd.Google Scholar
Furness, J. B. and Clerc, N. (2000). Responses of afferent neurons to the contents of the digestive tract, and their relation to endocrine and immune responses. Prog. Brain Res., 122, 159–172.CrossRefGoogle ScholarPubMed
Furness, J. B. and Costa, M. (1980). Types of nerves in the enteric nervous system. Neuroscience, 5, 1–20.Google Scholar
Furness, J. B. and Costa, M. (1987). The Enteric Nervous System. London: Churchill Livingstone.Google Scholar
Furness, J. B., Morris, J. L., Gibbins, I. L. and Costa, M. (1989). Chemical coding of neurons and plurichemical transmission. Annu. Rev. Pharmacol. Toxicol., 29, 289–306.CrossRefGoogle ScholarPubMed
Furness, J. B., Bornstein, J. C., Murphy, R. and Pompolo, S. (1992). Roles of peptides in transmission in the enteric nervous system. Trends Neurosci., 15, 66–71.CrossRefGoogle ScholarPubMed
Furness, J. B., Kunze, W. A., Bertrand, P. P., Clerc, N. and Bornstein, J. C. (1998). Intrinsic primary afferent neurons of the intestine. Prog. Neurobiol., 54, 1–18.CrossRefGoogle ScholarPubMed
Furness, J. B., Clerc, N., Vogalis, F. and Stebbing, M. J. (2003a). The enteric nervous system and its extrinsic connections. In Texbook of Gastroenterology, ed. Yamada, T., Alpers, D. H., Laine, L., Owyang, C. and Powell, D. W.. Philadelphia: Lippincott Williams & Wilkins, pp. 12–34.Google Scholar
Furness, J. B., Stebbing, M. J., Kunze, W. A. A. and Clerc, N. (2003b). Sensory neurons of the gastrointestinal tract. In Textbook of Gastroenterology, ed. Yamada, T., Alpers, D. H., Laine, L., Owyang, C. and Powell, D. W.. Philadelphia: Lippincott Williams & Wilkins, pp. 34–47.Google Scholar
Furness, J. B., Jones, C., Nurgali, K. and Clerc, N. (2004). Intrinsic primary afferent neurons and nerve circuits within the intestine. Prog. Neurobiol., 72, 143–164.CrossRefGoogle ScholarPubMed
Gabella, G. (1976). Structure of the Autonomic Nervous System. London: Chapman and Hall.CrossRefGoogle Scholar
Gamlin, P. D. R. (2000). Functions of the Edinger–Westphal nucleus. In The Autonomic Nervous System, Vol. 13, Nervous Control of the Eye, ed. Burnstock, G. and Sillito, A. M.. Amsterdam: Harwood Academic Publisher, pp. 117–154.Google Scholar
Gamlin, P. D. R. and Clarke, R. J. (1995). Single-unit activity in the primate nucleus reticularis tegmenti pontis related to vergence and ocular accommodation. J. Neurophysiol., 73, 2115–2119.CrossRefGoogle ScholarPubMed
Gamlin, P. D. R., Zhang, Y., Clendaniel, R. A. and Mays, L. E. (1994). Behavior of identified Edinger–Westphal neurons during ocular accommodation. J. Neurophysiol., 72, 2368–2382.CrossRefGoogle ScholarPubMed
Gandevia, S. C., Killian, K., McKenzie, D. K.et al. (1993). Respiratory sensations, cardiovascular control, kinaesthesia and transcranial stimulation during paralysis in humans. J. Physiol., 470, 85–107.CrossRefGoogle ScholarPubMed
Garrett, J. R., Ekström, J. and Anderson, L. C. (eds.) (1999). Frontiers in Oral Biology, Vol. 11, Basel: Karger.Google Scholar
Gaskell, W. H. (1916). The Involuntary Nervous System. London: Longmans.Google Scholar
Gauriau, C. and Bernard, J. F. (2002). Pain pathways and parabrachial circuits in the rat. Exp. Physiol., 87, 251–258.CrossRefGoogle ScholarPubMed
Gauriau, C. and Bernard, J. F. (2004a). Posterior triangular thalamic neurons convey nociceptive messages to the secondary somatosensory and insular cortices in the rat. J. Neurosci., 24, 752–761.CrossRefGoogle Scholar
Gauriau, C. and Bernard, J. F. (2004b). A comparative reappraisal of projections from the superficial laminae of the dorsal horn in the rat: the forebrain. J. Comp. Neurol., 468, 24–56.CrossRefGoogle Scholar
Gebhart, G. F. (ed.) (1995). Visceral Pain. Progress in Pain Research and Management. Seattle: IASP Press.Google Scholar
Gebhart, G. F. (1996). Visceral polymodal receptors. Prog. Brain Res., 113, 101–112.CrossRefGoogle ScholarPubMed
Gebhart, G. F. and Randich, A. (1992). Vagal modulation of nociception. Am. Pain Soc. J., 1, 26–32.Google Scholar
Gebber, G. L. (1990). Central determinants of sympathetic nerve discharge. In Central Regulation of Autonomic Functions, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 126–144.Google Scholar
Geerling, J. C., Mettenleiter, T. C. and Loewy, A. D. (2003). Orexin neurons project to diverse sympathetic outflow systems. Neuroscience, 122, 541–550.CrossRefGoogle ScholarPubMed
Gerber, U. and Polosa, C. (1978). Effects of pulmonary stretch receptor afferent stimulation on sympathetic preganglionic neuron firing. Can. J. Physiol. Pharmacol., 56, 191–198.CrossRefGoogle ScholarPubMed
Gerfen, C. R. and Sawchenko, P. E. (1984). An anterograde neuroanatomical tracing method that shows the detailed morphology of neurons, their axons and terminals: immunohistochemical localization of an axonally transported plant lectin, Phaseolus vulgaris leucoagglutinin (phaseolus vulgaris leuco-agglutinin). Brain Res., 290, 219–238.CrossRefGoogle Scholar
Gershon, M. D. (1994). Functional anatomy of the enteric nervous system. In Physiology of the Gastrointestinal Tract, 3rd edn., ed. Johnson, L. R.. New York: Raven Press, pp. 381–422.Google Scholar
Giamberardino, M. A. (1999). Recent and forgotten aspects of visceral pain. Eur. J. Pain, 3, 77–92.CrossRefGoogle ScholarPubMed
Gibbins, I. L. (1990). Target-related patterns of co-existence of neuropeptide Y, vasoactive intestinal peptide, enkephalin and substance P in cranial parasympathetic neurons innervating the facial skin and glands of guinea-pigs. Neuroscience, 38, 541–560.CrossRefGoogle ScholarPubMed
Gibbins, I. L. (1991). Vasomotor, pilomotor and secretomotor neurons distinguished by size and neuropeptide content in superior cervical ganglia of mice. J. Auton. Nerv. Syst., 34, 171–183.CrossRefGoogle Scholar
Gibbins, I. L. (1992). Vasoconstrictor, vasodilator and pilomotor pathways in sympathetic ganglia of guinea-pigs. Neuroscience, 47, 657–672.CrossRefGoogle ScholarPubMed
Gibbins, I. L. (1994). Comparative anatomy and evolution of the autonomic nervous system. In The Autonomic Nervous System, Vol. 4, Comparative Physiology and Evolution of the Autonomic Nervous System, ed. Nilsson, S. and Holmgren, S.. Chur: Harwood Academic Publishers, pp. 1–67.Google Scholar
Gibbins, I. L. (1995). Chemical neuroanatomy of sympathetic ganglia. In The Autonomic Nervous System, Vol. 6, Autonomic Ganglia, ed. McLachlan, E. M.. Luxembourg: Harwood Academic Publishers, pp. 73–122.Google Scholar
Gibbins, I. L. (1997). Autonomic pathways to cutaneous effectors. In The Autonomic Nervous System, Vol. 12, Autonomic Innervation of the Skin, ed. Morris, J. L. and Gibbins, I. L.. Chur, Switzerland: Harwood Academic Publishers, pp. 1–56.Google Scholar
Gibbins, I. L. (2004). Peripheral autonomic pathways. In The Human Nervous System, 2nd edn., ed. Paxinos, G. and Mai, J. K.. Amsterdam, San Diego, London: Elsevier Academic Press, pp. 134–189.Google Scholar
Gibbins, I. L. and Morris, J. L. (1987). Co-existence of neuropeptides in sympathetic, cranial autonomic and sensory neurons innervating the iris of the guinea-pig. J. Auton. Nerv. Syst., 21, 67–82.CrossRefGoogle ScholarPubMed
Gibbins, I. L. and Morris, J. L. (1990). Sympathetic noradrenergic neurons containing dynorphin but not neuropeptide Y innervate small cutaneous blood vessels of guinea-pigs. J. Auton. Nerv. Syst., 29, 137–149.CrossRefGoogle Scholar
Gibbins, I. L., Rodgers, H. F., Matthew, S. E. and Murphy, S. M. (1998). Synaptic organisation of lumbar sympathetic ganglia of guinea pigs: serial section ultrastructural analysis of dye-filled sympathetic final motor neurons. J. Comp. Neurol., 402, 285–302.3.0.CO;2-A>CrossRefGoogle ScholarPubMed
Gibbins, I. L., Jobling, P., Messenger, J. P., Teo, E. H. and Morris, J. L. (2000). Neuronal morphology and the synaptic organisation of sympathetic ganglia. J. Auton. Nerv. Syst., 81, 104–109.CrossRefGoogle ScholarPubMed
Gibbins, I. L., Jobling, P. and Morris, J. L. (2003a). Functional organization of peripheral vasomotor pathways. Acta Physiol. Scand., 177, 237–245.CrossRefGoogle Scholar
Gibbins, I. L., Jobling, P., Teo, E. H., Matthew, S. E. and Morris, J. L. (2003b). Heterogeneous expression of SNAP-25 and synaptic vesicle proteins by central and peripheral inputs to sympathetic neurons. J. Comp. Neurol., 459, 25–43.CrossRefGoogle Scholar
Gibbins, I. L., Teo, E. H., Jobling, P. and Morris, J. L. (2003c). Synaptic density, convergence, and dendritic complexity of prevertebral sympathetic neurons. J. Comp. Neurol., 455, 285–298.CrossRefGoogle Scholar
Gilbey, M. P., Jordan, D., Richter, D. W. and Spyer, K. M. (1984). Synaptic mechanisms involved in the inspiratory modulation of vagal cardio-inhibitory neurones in the cat. J. Physiol., 356, 65–78.CrossRefGoogle ScholarPubMed
Gilliat, R. W. (1948). Vaso-constriction in the finger after deep inspiration. J. Physiol., 107, 76–88.CrossRefGoogle Scholar
Gilliat, R. W., Guttmann, L. and Whitteridge, D. (1948). Inspiratory vasoconstriction in patients after spinal injuries. J. Physiol., 107, 67–75.CrossRefGoogle Scholar
Giuliano, F., Allard, J., Compagnie, S.et al. (2001). Vaginal physiological changes in a model of sexual arousal in anesthetized rats. Am. J. Physiol. Regul. Integr. Comp. Physiol., 281, R140–R149.CrossRefGoogle Scholar
Gladwell, S. J. and Coote, J. H. (1999). Inhibitory and indirect excitatory effects of dopamine on sympathetic preganglionic neurones in the neonatal rat spinal cord in vitro. Brain Res., 818, 397–407.CrossRefGoogle ScholarPubMed
Goehler, L. E., Gaykema, R. P., Hansen, M. K.et al. (2000). Vagal immune-to-brain communication: a visceral chemosensory pathway. Auton. Neurosci., 85, 49–59.CrossRefGoogle ScholarPubMed
Gola, M. and Niel, J. P. (1993). Electrical and integrative properties of rabbit sympathetic neurones re-evaluated by patch clamping non-dissociated cells. J. Physiol., 460, 327–349.CrossRefGoogle ScholarPubMed
Goldstein, D. S. (1995). Stress, Catecholamines, and Cardiovascular Disease. New York, Oxford: Oxford University Press.Google Scholar
Goldstein, D. S. (2000). The Autonomic Nervous System in Health and Disease. New York: Marcel Dekker.Google Scholar
Golenhofen, K., Hensel, H. and Ruef, J. (1962). Sustained dilatation in human muscle blood vessels under the influence of adrenaline. J. Physiol., 160, 189–199.CrossRefGoogle ScholarPubMed
Goodchild, A. K., Moon, E. A., Dampney, R. A. and Howe, P. R. (1984). Evidence that adrenaline neurons in the rostral ventrolateral medulla have a vasopressor function. Neurosci. Lett., 45, 267–272.CrossRefGoogle ScholarPubMed
Goodchild, A. K., Deurzen, B. T. M., Sun, Q. J., Chalmers, J. and Pilowsky, P. (2000). Spinal γ-aminobutyric acid receptors do not mediate the sympathetic baroreceptor reflex in the rat. Am. J. Physiol. Regul. Integr. Comp. Physiol., 279, R320–R331.CrossRefGoogle Scholar
Goodwin, G. M., McCloskey, D. I. and Mitchell, J. H. (1972). Cardiovascular and respiratory responses to changes in central command during isometric exercise at constant muscle tension. J. Physiol., 226, 173–190.CrossRefGoogle ScholarPubMed
Gordon, F. J. and McCann, L. A. (1988). Pressor responses evoked by microinjections of L-glutamate into the caudal ventrolateral medulla of the rat. Brain Res., 457, 251–258.CrossRefGoogle ScholarPubMed
Gore, A. C. and Roberts, J. L. (2003). Neuroendocrine systems. In Fundamental Neuroscience, 2nd edn., ed. Squire, L. R., Bloom, F. E., McConnell, S. K.et al. San Diego: Academic Press, pp. 1031–1065.Google Scholar
Goridis, C. and Rohrer, H. (2002). Specification of catecholaminergic and serotonergic neurons. Nat. Rev. Neurosci., 3, 531–541.CrossRefGoogle ScholarPubMed
Gould, D. J. and Hill, C. E. (1996). Alpha-adrenoceptor activation of a chloride conductance in rat iris arterioles. Am. J. Physiol., 271, H2469–H2476.Google ScholarPubMed
Granit, R. (1981). Comments on history of motor control. In Handbook of Physiology, Section I, The Nervous System, Vol. II, Motor Control, part I. ed. Brooks, V. B.. Bethesda: American Physiological Society, pp. 1–16.Google Scholar
Grasby, D. J., Gibbins, I. L. and Morris, J. L. (1997). Projections of sympathetic non-noradrenergic neurons to skeletal muscle arteries in guinea-pig limbs vary with the metabolic character of muscles. J. Vasc. Res., 34, 351–364.CrossRefGoogle ScholarPubMed
Gray's Anatomy (1995). 38th edn. Edinburgh: Churchill Livingstone.
Graziano, A. and Jones, E. G. (2004). Widespread thalamic terminations of fibers arising in the superficial medullary dorsal horn of monkeys and their relation to calbindin immunoreactivity. J. Neurosci., 24, 248–256.CrossRefGoogle ScholarPubMed
Green, P. G., Miao, F. J. P., Jänig, W. and Levine, J. D. (1995). Negative feedback neuroendocrine control of the inflammatory response in rats. J. Neurosci., 15, 4678–4686.CrossRefGoogle ScholarPubMed
Green, P. G., Jänig, W. and Levine, J. D. (1997). Negative feedback neuroendocrine control of inflammatory response in the rat is dependent on the sympathetic postganglionic neuron. J. Neurosci., 17, 3234–3238.CrossRefGoogle ScholarPubMed
Greenfield, A. D. (1966). Survey of the evidence for active neurogenic vasodilation in man. Fed. Proc., 25, 1607–1610.Google Scholar
Greger, R. and Windhorst, U. (eds.) (1996). Comprehensive Human Physiology. From Cellular Mechanisms to Integration, Vol. 1 and 2. Berlin, Heidelberg, New York: Springer Verlag.CrossRefGoogle Scholar
Gregor, M. and Jänig, W. (1977). Effects of systemic hypoxia and hypercapnia on cutaneous and muscle vasoconstrictor neurones to the cat's hindlimb. Pflügers Arch., 368, 71–81.CrossRefGoogle ScholarPubMed
Gregor, M., Jänig, W. and Riedel, W. (1976). Response pattern of cutaneous postganglionic neurones to the hindlimb on spinal cord heating and cooling in the cat. Pflügers Arch., 363, 135–140.CrossRefGoogle ScholarPubMed
Grewe, W., Jänig, W. and Kümmel, H. (1995). Effects of hypothalamic thermal stimuli on sympathetic neurones innervating skin and skeletal muscle of the cat hindlimb. J. Physiol., 488, 139–152.CrossRefGoogle ScholarPubMed
Griffith, W. H. III, Gallagher, J. P. and Shinnick-Gallagher, P. (1980). An intracellular investigation of cat vesical pelvic ganglia. J. Neurophysiol., 43, 343–354.CrossRefGoogle ScholarPubMed
Grigg, P., Schaible, H. G. and Schmidt, R. F. (1986). Mechanical sensitivity of group III and IV afferents from posterior articular nerve in normal and inflamed cat knee. J. Neurophysiol., 55, 635–643.CrossRefGoogle ScholarPubMed
Grosse, M. and Jänig, W. (1976). Vasoconstrictor and pilomotor fibres in skin nerves to the cat's tail. Pflügers Arch., 361, 221–229.CrossRefGoogle ScholarPubMed
Grubb, B. P. and Karas, B. (2002). Neurally mediated syncope. In Autonomic Failure, 4th edn., ed. Mathias, C. J. and Bannister, R.. Oxford: Oxford University Press, pp. 437–447.Google ScholarPubMed
Grundy, D. (1988). Speculation on the structure/function relationship for vagal and splanchnic afferent endings supplying the gastrointestinal tract. J. Auton. Nerv. Syst., 22, 175–180.CrossRefGoogle Scholar
Grundy, D. and Scratcherd, T. (1989). Sensory afferents from the gastrointestinal tract. In Handbook of Physiology, Section 6: The Gastrointestinal System, Vol. 1: Motility and Circulation, ed. Wood, J. D.. Bethesda: American Physiological Society, pp. 593–620.Google Scholar
Grundy, D., Salih, A. A. and Scratcherd, T. (1981). Modulation of vagal efferent fibre discharge by mechanoreceptors in the stomach, duodenum and colon of the ferret. J. Physiol., 319, 43–52.CrossRefGoogle ScholarPubMed
Guertzenstein, P. G. and Silver, A. (1974). Fall in blood pressure produced from discrete regions of the ventral surface of the medulla by glycine and lesions. J. Physiol., 242, 489–503.CrossRefGoogle ScholarPubMed
Guth, L. and Bernstein, J. J. (1961). Selectivity in the re-establishment of synapses in the superior cervical sympathetic ganglion of the cat. Exp. Neurol., 4, 59–69.CrossRefGoogle ScholarPubMed
Guttmann, L. (1976). Spinal Cord Injuries. Comprehensive Management and Research. Oxford: Blackwell Scientific Publications.Google Scholar
Guyenet, P. G. (1990). Role of the ventral medulla oblongata in blood pressure regulation. In Central Regulation of Autonomic Functions, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 145–167.Google Scholar
Guyenet, P. G. (2000). Neural structures that mediate sympathoexcitation during hypoxia. Respir. Physiol., 121, 147–162.CrossRefGoogle ScholarPubMed
Guyenet, P. G. and Brown, D. L. (1986). Nucleus paragigantocellularis lateralis and lumbar sympathetic discharge in the rat. Am. J. Physiol., 250, R1081–R1094.Google ScholarPubMed
Guyenet, P. G. and Koshiya, N. (1992). Respiratory-sympathetic integration in the medulla oblongata. In Central Neural Mechanisms in Cardiovascular Regulation, ed. Kunos, G. and Ciriello, J.. Boston: Birkhäuser, pp. 226–247.CrossRefGoogle Scholar
Guyenet, P. G. and Stornetta, R. L. (1997). Central nervous system regulation of the sympathetic and cardiovagal vasomotor outflows. In Anesthesia, ed. Yaksh, T. L., Maze, M., Lynch, C.et al. Philadelphia, New York: Lippincott-Raven, pp. 1205–1232.Google Scholar
Guyenet, P. G., Koshiya, N., Huangfu, D.et al. (1996). Role of medulla oblongata in generation of sympathetic and vagal outflows. Prog. Brain Res., 107, 127–144.CrossRefGoogle ScholarPubMed
Guyenet, P. G., Schreihofer, A. M. and Stornetta, R. L. (2001). Regulation of sympathetic tone and arterial pressure by the rostral ventrolateral medulla after depletion of C1 cells in rats. Ann. New York Acad. Sci., 940, 259–269.CrossRefGoogle ScholarPubMed
Häbler, H. J. and Jänig, W. (1995). Coordination of sympathetic and respiratory systems: neurophysiological experiments. Clin. Exp. Hypertens., 17, 223–235.CrossRefGoogle ScholarPubMed
Häbler, H. J., Jänig, W. and Koltzenburg, M. (1990a). Activation of unmyelinated afferent fibres by mechanical stimuli and inflammation of the urinary bladder in the cat. J. Physiol., 425, 545–562.CrossRefGoogle Scholar
Häbler, H. J., Jänig, W., Koltzenburg, M. and McMahon, S. B. (1990b). A quantitative study of the central projection patterns of unmyelinated ventral root afferents in the cat. J. Physiol., 422, 265–287.CrossRefGoogle Scholar
Häbler, H. J., Hilbers, K., Jänig, W.et al. (1992). Viscero-sympathetic reflex responses to mechanical stimulation of pelvic viscera in the cat. J. Auton. Nerv. Syst., 38, 147–158.CrossRefGoogle ScholarPubMed
Häbler, H. J., Jänig, W. and Koltzenburg, M. (1993a). Myelinated primary afferents of the sacral spinal cord responding to slow filling and distension of the urinary bladder. J. Physiol., 463, 449–460.CrossRefGoogle Scholar
Häbler, H. J., Jänig, W. and Koltzenburg, M. (1993b). Receptive properties of myelinated primary afferents innervating the inflamed urinary bladder of the cat. J. Neurophysiol., 69, 395–405.CrossRefGoogle Scholar
Häbler, H. J., Jänig, W., Krummel, M. and Peters, O. A. (1993c). Respiratory modulation of the activity in postganglionic neurons supplying skeletal muscle and skin of the rat hindlimb. J. Neurophysiol., 70, 920–930.CrossRefGoogle Scholar
Häbler, H. J., Jänig, W., Krummel, M. and Peters, O. A. (1994a). Reflex patterns in postganglionic neurons supplying skin and skeletal muscle of the rat hindlimb. J. Neurophysiol., 72, 2222–2236.CrossRefGoogle Scholar
Häbler, H. J., Jänig, W. and Michaelis, M. (1994b). Respiratory modulation of activity in sympathetic neurones. Prog. Neurobiol., 43, 567–606.CrossRefGoogle Scholar
Häbler, H. J., Bartsch, T. and Jänig, W. (1996). Two distinct mechanisms generate the respiratory modulation in fibre activity of the rat cervical sympathetic trunk. J. Auton. Nerv. Syst., 61, 116–122.CrossRefGoogle ScholarPubMed
Häbler, H. J., Boczek-Funcke, A., Michaelis, M. and Jänig, W. (1997a). Responses of distinct types of sympathetic neuron to stimulation of the superior laryngeal nerve in the cat. J. Auton. Nerv. Syst., 66, 97–104.CrossRefGoogle Scholar
Häbler, H. J., Wasner, G., Bartsch, T. and Jänig, W. (1997b). Responses of rat postganglionic sympathetic vasoconstrictor neurons following blockade of nitric oxide synthesis in vivo. Neuroscience, 77, 899–909.CrossRefGoogle Scholar
Häbler, H. J., Wasner, G. and Jänig, W. (1997c). Interaction of sympathetic vasoconstriction and antidromic vasodilatation in the control of skin blood flow. Exp. Brain Res., 113, 402–410.CrossRefGoogle Scholar
Häbler, H. J., Bartsch, T. and Jänig, W. (1999a). Rhythmicity in single fiber postganglionic activity supplying the rat tail. J. Neurophysiol., 81, 2026–2036.CrossRefGoogle Scholar
Häbler, H. J., Timmermann, L., Stegmann, J. U. and Jänig, W. (1999b). Involvement of neurokinins in antidromic vasodilatation in hairy and hairless skin of the rat hindlimb. Neuroscience, 89, 1259–1268.CrossRefGoogle Scholar
Häbler, H. J., Bartsch, T. and Jänig, W. (2000). Respiratory rhythmicity in the activity of postganglionic neurones supplying the rat tail during hyperthermia. J. Auton. Nerv. Syst., 83, 75–80.Google ScholarPubMed
Hadziefendic, S. and Haxhiu, M. A. (1999). central nervous system innervation of vagal preganglionic neurons controlling peripheral airways: a transneuronal labeling study using pseudorabies virus. J. Auton. Nerv. Syst., 76, 135–145.CrossRefGoogle ScholarPubMed
Hagbarth, K. E. and Vallbo, A. B. (1968). Pulse and respiratory grouping of sympathetic impulses in human muscle nerves. Acta Physiol. Scand., 74, 96–108.CrossRefGoogle ScholarPubMed
Hagbarth, K. E., Hallin, R. G., Hongell, A., Torebjörk, H. E. and Wallin, B. G. (1972). General characteristics of sympathetic activity in human skin nerves. Acta Physiol. Scand., 84, 164–176.CrossRefGoogle ScholarPubMed
Hainsworth, R. (2002). Syncope and fainting: classification pathophysiological basis. In Autonomic Failure, 4th edn., eds. Mathias, C. J. and Bannister, R.. Oxford: Oxford University Press, pp. 428–436.Google ScholarPubMed
Hall, M. (1841). On the Diseases and Derangements of the Nervous System, in their Primary Forms and their Modifications by Age, Sex Constitution, Heredity, Disposition, Excesses, General Disorder, and Organic Disease. London: Bailliere.Google Scholar
Halliday, G. M. and McLachlan, E. M. (1991). A comparative analysis of neurons containing catecholamine-synthesizing enzymes and neuropeptide Y in the ventrolateral medulla of rats, guinea-pigs and cats. Neuroscience, 43, 531–550.CrossRefGoogle Scholar
Hallin, R. G. and Torebjörk, H. E. (1974). Single unit sympathetic activity in human skin nerves during rest and various manoeuvres. Acta Physiol. Scand., 92, 303–317.CrossRefGoogle ScholarPubMed
Hamblin, P. A., McLachlan, E. M. and Lewis, R. J. (1995). Sub-nanomolar concentrations of ciguatoxin-1 excite preganglionic terminals in guinea pig sympathetic ganglia. Naunyn-Schmiedeberg's Arch. Pharmacol., 352, 236–246.CrossRefGoogle ScholarPubMed
Harden, R. N., Baron, R. and Jänig, W. (eds.) (2001). Complex Regional Pain Syndrome. Seattle: IASP Press.Google ScholarPubMed
Hargreaves, K. M., Roszkowski, M. T. and Swift, J. Q. (1993). Bradykinin and inflammatory pain. Agents Actions Suppl., 41, 65–73.Google ScholarPubMed
Harhun, M. I., Pucovsky, V., Povstyan, O. V., Gordienko, D. V. and Bolton, T. B. (2005). Interstitial cells in the vasculature. J. Cell Mol. Med., 9, 232–243.CrossRefGoogle ScholarPubMed
Haselton, J. R. and Guyenet, P. G. (1989). Central respiratory modulation of medullary sympathoexcitatory neurons in rat. Am. J. Physiol., 256, R739–R750.Google ScholarPubMed
Haupt, P., Jänig, W. and Kohler, W. (1983). Response pattern of visceral afferent fibres, supplying the colon, upon chemical and mechanical stimuli. Pflügers Arch., 398, 41–47.CrossRefGoogle ScholarPubMed
Haxhiu, M. A. and Loewy, A. D. (1996). Central connections of the motor and sensory vagal systems innervating the trachea. J. Auton. Nerv. Syst., 57, 49–56.CrossRefGoogle ScholarPubMed
Haxhiu, M. A., Jansen, A. S., Cherniack, N. S. and Loewy, A. D. (1993). central nervous system innervation of airway-related parasympathetic preganglionic neurons: a transneuronal labeling study using pseudorabies virus. Brain Res., 618, 115–134.CrossRefGoogle ScholarPubMed
Haxhiu, M. A., Erokwu, B., Bhardwaj, V. and Dreshaj, I. A. (1998). The role of the medullary raphe nuclei in regulation of cholinergic outflow to the airways. J. Auton. Nerv. Syst., 69, 64–71.CrossRefGoogle ScholarPubMed
Heel, K. A., McCauley, R. D., Papadimitriou, J. M. and Hall, J. C. (1997). Review: Peyer's patches. J. Gastroenterol. Hepatol., 12, 122–136.CrossRefGoogle ScholarPubMed
Hellmann, K. (1963). The effect of temperature changes on the isolated pilomotor muscles. J. Physiol., 169, 621–629.CrossRefGoogle ScholarPubMed
Henderson, C. G. and Ungar, A. (1978). Effect of cholinergic antagonists on sympathetic ganglionic transmission of vasomotor reflexes from the carotid baroreceptors and chemoreceptors of the dog. J. Physiol., 277, 379–385.CrossRefGoogle ScholarPubMed
Henderson, L. A., Keay, K. A. and Bandler, R. (1998). The ventrolateral periaqueductal gray projects to caudal brainstem depressor regions: a functional-anatomical and physiological study. Neuroscience, 82, 201–221.CrossRefGoogle ScholarPubMed
Hendry, I. A. and Hill, C. E. (eds.) (1992). The Autonomic Nervous System, Vol. 2, Development, Regeneration and Plasticity, Chur: Harwood Academic Publishers.Google Scholar
Henry, J. P. (1997). Culture and High Blood Pressure. Hamburg, Münster: LIT Publishing Company.Google Scholar
Henry, J. P. and Grim, C. E. (1990). Psychosocial mechanisms of primary hypertension. J. Hypertens., 8, 783–793.CrossRefGoogle ScholarPubMed
Henry, J. P. and Stephens, P. M. (1977). Stress, Health and the Social Environment: a Sociobiologic Approach to Medicine. Heidelberg, Berlin: Springer.CrossRefGoogle Scholar
Hensel, H. (1981). Thermoreception and Temperature Regulation. London, New York: Academic Press.Google ScholarPubMed
Hensel, H. (1982). Thermal Sensations and Thermoreceptors in Man. Springfield Illinois: Charles C. Thomas Publ.Google Scholar
Herbert, H., Moga, M. M. and Saper, C. B. (1990). Connections of the parabrachial nucleus with the nucleus of the solitary tract and the medullary reticular formation in the rat. J. Comp. Neurol., 293, 540–580.CrossRefGoogle ScholarPubMed
Hering, E. (1869). Über den Einfluβ der Atmung auf den Kreislauf. Erste Mitteilung. Über Atembewegungen der Gefäβ systeme [On the influence of the respiration on the circulation system. First report. On the respiratory movement of the vascular systems]. Sitzungsber. Akad. Wiss. Wien, Math.- Naturw., Abt. 2, 60, 829–856.Google Scholar
Hering, H. E. (1927). Die Karotissinusreflexe [The Carotid Sinus Reflexes]. Dresden, Leipzig: Verlag von Theodor Steinkopf.Google Scholar
Hermann, G. and Rogers, R. C. (1995). Tumor necrosis factor-alpha in the dorsal vagal complex suppresses gastric motility. NeuroImmunoModulation, 2, 74–81.CrossRefGoogle ScholarPubMed
Hermann, G. E., Tovar, C. A. and Rogers, R. C. (2002). LPS-induced suppression of gastric motility relieved by tumor necrosis factorR:Fc construct in dorsal vagal complex. Am. J. Physiol. Gastrointest. Liver Physiol., 283, G634–G639.CrossRefGoogle Scholar
Hertz, A. F. (1911). The Sensibility of the Alimentary Canal. Oxford: Oxford University Press.Google Scholar
Hess, W. R. (1944). Hypothalamische Adynamie [Hypothalamic adynamia]. Helv. Physiol. Acta, 2, 137–147.Google Scholar
Hess, W. R. (1948). Die Organisation des vegetativen Nervensystems [The Organization of the Autonomic Nervous System]. Basel: Benno Schwabe & Co.Google Scholar
Hess, W. R. (1954). Das Zwischenhirn, Syndrome, Lokalisationen, Funktionen [The Diencephalon, Syndromes, Localizations, Functions], 2nd edn. Basel: Benno Schwabe.Google Scholar
Hess, W. R. and Brügger, M. (1943). Das subcortikale Zentrum der affektiven Abwehrreaktion [The subcortical center of the affective defense reaction]. Helv. Physiol. Acta, 1, 33–52.Google Scholar
Heuckeroth, R. O., Enomoto, H., Grider, J. R.et al. (1999). Gene targeting reveals a critical role for neurturin in the development and maintenance of enteric, sensory, and parasympathetic neurons. Neuron, 22, 253–263.CrossRefGoogle ScholarPubMed
Heymans, C. and Neil, E. (1958). Reflexogenic Areas of the Cardiovascular System. London: J. & A. Churchill Ltd.Google Scholar
Hill, C. E., Hendry, I. A. and Sheppard, A. (1987). Use of the fluorescent dye, fast blue, to label sympathetic postganglionic neurones supplying mesenteric arteries and enteric neurones of the rat. J. Auton. Nerv. Syst., 18, 73–82.CrossRefGoogle ScholarPubMed
Hill, C. E., Klemm, M., Edwards, F. R. and Hirst, G. D. (1993). Sympathetic transmission to the dilator muscle of the rat iris. J. Auton. Nerv. Syst., 45, 107–123.CrossRefGoogle ScholarPubMed
Hill, C. E., Eade, J. and Sandow, S. L. (1999). Mechanisms underlying spontaneous rhythmical contractions in irideal arterioles of the rat. J. Physiol., 521, 507–516.CrossRefGoogle ScholarPubMed
Hill, M. A., Zou, H., Potocnik, S. J., Meininger, G. A. and Davis, M. J. (2001). Invited review: arteriolar smooth muscle mechanotransduction: Ca2 + signaling pathways underlying myogenic reactivity. J. Appl. Physiol., 91, 973–983.CrossRefGoogle Scholar
Hillsley, K. and Grundy, D. (1998). Sensitivity to 5-hydroxytryptamine in different afferent subpopulations within mesenteric nerves supplying the rat jejunum. J. Physiol., 509, 717–727.CrossRefGoogle ScholarPubMed
Hillsley, K., Kirkup, A. J. and Grundy, D. (1998). Direct and indirect actions of 5-hydroxytryptamine on the discharge of mesenteric afferent fibres innervating the rat jejunum. J. Physiol., 506, 551–561.CrossRefGoogle ScholarPubMed
Himms-Hagen, J. (1991). Neural control of brown adipose tissue: thermogenesis, hypertrophy, and atrophy neuroeffector junctions. Front. Neuroendocrinol., 12, 38–93.Google Scholar
Hirooka, Y., Polson, J. W., Potts, P. D. and Dampney, R. A. (1997). Hypoxia-induced Fos expression in neurons projecting to the pressor region in the rostral ventrolateral medulla. Neuroscience, 80, 1209–1224.CrossRefGoogle ScholarPubMed
Hirshberg, R. M., Al Chaer, E. D., Lawand, N. B., Westlund, K. N. and Willis, W. D. (1996). Is there a pathway in the posterior funiculus that signals visceral pain?Pain, 67, 291–305.CrossRefGoogle Scholar
Hirst, G. D. (2001). An additional role for interstitial cell of Cajal in the control of gastrointestinal motility?J. Physiol., 537, 1CrossRefGoogle Scholar
Hirst, G. D. and Edwards, F. R. (1989). Sympathetic neuroeffector transmission in arteries and arterioles. Physiol. Rev., 69, 546–604.CrossRefGoogle ScholarPubMed
Hirst, G. D. and McLachlan, E. M. (1984). Post-natal development of ganglia in the lower lumbar sympathetic chain of the rat. J. Physiol., 349, 119–134.CrossRefGoogle ScholarPubMed
Hirst, G. D. and McLachlan, E. M. (1986). Development of dendritic calcium currents in ganglion cells of the rat lower lumbar sympathetic chain. J. Physiol., 377, 349–368.CrossRefGoogle ScholarPubMed
Hirst, G. D. and Neild, T. O. (1980). Some properties of spontaneous excitatory junction potentials recorded from arterioles of guinea-pigs. J. Physiol., 303, 43–60.CrossRefGoogle ScholarPubMed
Hirst, G. D. and Ward, S. M. (2003). Interstitial cells: involvement in rhythmicity and neural control of gut smooth muscle. J. Physiol., 550, 337–346.CrossRefGoogle ScholarPubMed
Hirst, G. D. S., Holman, M. E. and McKirdy, H. C. (1974). Two types of neurones in the myenteric plexus of duodenum in the guinea-pig. J. Physiol., 236, 303–326.CrossRefGoogle ScholarPubMed
Hirst, G. D., Holman, M. E. and McKirdy, H. C. (1975). Two descending nerve pathways activated by distension of guinea-pig small intestine. J. Physiol., 244, 113–127.CrossRefGoogle ScholarPubMed
Hirst, G. D. S., Edwards, F. R., Bramich, N. J. and Klemm, M. F. (1991). Neural control of cardiac pacemaker potentials. News Physiol. Sci., 6, 185–190.Google Scholar
Hirst, G. D. S., Bramich, N. J., Edwards, F. R. and Klemm, M. (1992). Transmission at autonomic neuroeffector junctions. Trends Neurosci., 15, 40–46.Google ScholarPubMed
Hirst, G. D., Choate, J. K., Cousins, H. M., Edwards, F. R. and Klemm, M. F. (1996). Transmission by post-ganglionic axons of the autonomic nervous system: the importance of the specialized neuroeffector junction. Neuroscience, 73, 7–23.CrossRefGoogle ScholarPubMed
Hoffmeister, B., Hussels, W. and Jänig, W. (1978). Long-lasting discharge of postganglionic neurones to skin and muscle of the cat's hindlimb after repetitive activation of preganglionic axons in the lumbar sympathetic trunk. Pflügers Arch., 376, 15–20.CrossRefGoogle ScholarPubMed
Hohmann, E. L., Elde, R. P., Rysavy, J. A., Einzig, S. and Gebhard, R. L. (1986). Innervation of periosteum and bone by sympathetic vasoactive intestinal peptide-containing nerve fibers. Science, 232, 868–871.CrossRefGoogle ScholarPubMed
Holman, M. E., Coleman, H. A., Tonta, M. A. and Parkington, H. C. (1994). Synaptic transmission from splanchnic nerves to the adrenal medulla of guinea-pigs. J. Physiol., 478, 115–124.CrossRefGoogle ScholarPubMed
Holzer, P. (1992). Peptidergic sensory neurons in the control of vascular functions: mechanisms and significance in the cutaneous and splanchnic vascular beds. Rev. Physiol. Biochem. Pharmacol., 121, 49–146.CrossRefGoogle ScholarPubMed
Holzer, P. (1995). Chemosensitive afferent nerves in the regulation of gastric blood flow and protection. Adv. Exp. Med. Biol., 371B, 891–895.Google ScholarPubMed
Holzer, P. (1998a). Neural emergency system in the stomach. Gastroenterology, 114, 823–839.CrossRefGoogle Scholar
Holzer, P. (1998b). Neurogenic vasodilatation and plasma leakage in the skin. Gen. Pharmacol., 30, 5–11.CrossRefGoogle Scholar
Holzer, P. (2002a). Sensory neurone responses to mucosal noxae in the upper gut: relevance to mucosal integrity and gastrointestinal pain. Neurogastroenterol. Motil., 14, 459–475.CrossRefGoogle Scholar
Holzer, P. (2002b). Control of gastric functions by extrinsic sensory neurons. In The Autonomic Nervous System, Vol. 14, Innervation of the Gastrointestinal Tract, ed. Brookes, S. and Costa, M., London, New York: Taylor and Francis, pp. 103–170.Google Scholar
Holzer, P. (2003). Afferent signalling of gastric acid challenge. J. Physiol. Pharmacol., 54 Suppl4, 43–53.Google ScholarPubMed
Holzer, P. and Maggi, C. A. (1998). Dissociation of dorsal root ganglion neurons into afferent and efferent-like neurons. Neuroscience, 86, 389–398.Google ScholarPubMed
Hopkins, D. A., Bieger, D., deVente, J. and Steinbusch, W. M. (1996). Vagal efferent projections: viscerotopy, neurochemistry and effects of vagotomy. Prog. Brain Res., 107, 79–96.CrossRefGoogle ScholarPubMed
Horeyseck, G. and Jänig, W. (1974a). Reflexes in postganglionic fibres within skin and muscle nerves after mechanical non-noxious stimulation of skin. Exp. Brain Res., 20, 115–123.Google Scholar
Horeyseck, G. and Jänig, W. (1974b). Reflexes in postganglionic fibres within skin and muscle nerves after noxious stimulation of skin. Exp. Brain Res., 20, 125–134.Google Scholar
Horeyseck, G. and Jänig, W. (1974c). Reflex activity in postganglionic fibres within skin and muscle nerves elicited by somatic stimuli in chronic spinal cats. Exp. Brain Res., 21, 155–168.CrossRefGoogle Scholar
Horeyseck, G., Jänig, W., Kirchner, F. and Thämer, V. (1976). Activation and inhibition of muscle and cutaneous postganglionic neurones to hindlimb during hypothalamically induced vasoconstriction and atropine-sensitive vasodilation. Pflügers Arch., 361, 231–240.CrossRefGoogle ScholarPubMed
Hori, T., Katafuchi, T., Take, S., Shimizu, N. and Niijima, A. (1995). The autonomic nervous system as a communication channel between the brain and the immune system. NeuroImmunoModulation, 2, 203–215.CrossRefGoogle ScholarPubMed
Horiuchi, J. and Dampney, R. A. L. (2002). Evidence for tonic disinhibition of rostral ventrolateral medulla sympathoexcitatory neurons from the caudal pressor area. Auton. Neurosci., 99, 102–110.CrossRefGoogle ScholarPubMed
Horiuchi, J., Killinger, S. and Dampney, R. A. (2004a). Contribution to sympathetic vasomotor tone of tonic glutamatergic inputs to neurons in the rostral ventrolateral medulla. Am. J. Physiol. Regul. Integr. Comp. Physiol., 287, R1335–R1343.CrossRefGoogle Scholar
Horiuchi, J., McAllen, R. M., Allen, A. M.et al. (2004b). Descending vasomotor pathways from the dorsomedial hypothalamic nucleus: role of medullary raphe and rostral ventrolateral medulla. Am. J. Physiol. Regul. Integr. Comp. Physiol., 287, R824–R832.CrossRefGoogle Scholar
Horowitz, B., Ward, S. M. and Sanders, K. M. (1999). Cellular and molecular basis for electrical rhythmicity in gastrointestinal muscles. Annu. Rev. Physiol., 61, 19–43.CrossRefGoogle ScholarPubMed
Hosoya, Y., Sugiura, Y., Okado, N., Loewy, A. D. and Kohno, K. (1991). Descending input from the hypothalamic paraventricular nucleus to sympathetic preganglionic neurons in the rat. Exp. Brain Res., 85, 10–20.CrossRefGoogle ScholarPubMed
Hoyle, C. H. V., Milner, P. and Burnstock, G. (2002). Neuroeffector transmission in the intestine. In The Autonomic Nervous System, Vol. 14, Innervation of the Gastrointestinal Tract, ed. Brookes, S. and Costa, M.. London, New York: Taylor and Francis, pp. 295–340.Google Scholar
Hua, L. H., Strigo, I. A., Baxter, L. C., Johnson, S. C. and Craig, A. D. (2005). Anterior-posterior somatotopy of innocuous cooling activation focus in human dorsal insular cortex. Am. J. Physiol. Regul. Intrgr. Comp. Physiol., 289, R319–R325.CrossRefGoogle Scholar
Huang, J. and Weiss, M. L. (1999). Characterization of the central cell groups regulating the kidney in the rat. Brain Res., 845, 77–91.CrossRefGoogle ScholarPubMed
Huang, X. F., Törk, I. and Paxinos, G. (1993). Dorsal motor nucleus of the vagus nerve: a cyto- and chemoarchitectonic study in the human. J. Comp. Neurol., 330, 158–182.CrossRefGoogle ScholarPubMed
Huizinga, J. D., Thuneberg, L., Vanderwinden, J. M. and Rumessen, J. J. (1997). Interstitial cells of Cajal as targets for pharmacological intervention in gastrointestinal motor disorders. Trends Pharmacol. Sci., 18, 393–403.CrossRefGoogle ScholarPubMed
Huizinga, J. D., Ambrous, K. and Der-Silaphet, T. (1998). Co-operation between neural and myogenic mechanisms in the control of distension-induced peristalsis in the mouse small intestine. J. Physiol., 506, 843–856.CrossRefGoogle ScholarPubMed
Huizinga, J. D. and Faussone-Pellegrini, M. S. (2005). About the presence of interstitial cells of Cajal outside the musculature of the gastrointestinal tract. J. Cell Mol. Med., 9, 468–473.CrossRefGoogle ScholarPubMed
Hultborn, H. (2001). State-dependent modulation of sensory feedback. J. Physiol., 533, 5–13.CrossRefGoogle ScholarPubMed
Hummel, T., Sengupta, J. N., Meller, S. T. and Gebhart, G. F. (1997). Responses of T2–4 spinal cord neurons to irritation of the lower airways in the rat. Am. J. Physiol., 273, R1147–R1157.Google ScholarPubMed
Hunsperger, R. W. (1956). Affektreactionen auf elektrische Reizung im Hirnstamm der Katze [Affect reactions to electrical stimulation in the brain stem of the cat]. Helv. Physiol. Acta, 14, 70–92.Google Scholar
Iino, S., Ward, S. M. and Sanders, K. M. (2004). Interstitial cells of Cajal are functionally innervated by excitatory motor neurones in the murine intestine. J. Physiol., 556, 521–530.CrossRefGoogle ScholarPubMed
Inoue, T. (1980). Efferent discharge patterns in the ciliary nerve of rabbits and the pupillary light reflex. Brain Res., 186, 43–53.CrossRefGoogle ScholarPubMed
Ireland, D. R. (1999). Preferential formation of strong synapses during re-innervation of guinea-pig sympathetic ganglia. J. Physiol., 520, 827–837.CrossRefGoogle ScholarPubMed
Ireland, D. R., Davies, P. J. and McLachlan, E. M. (1999). Calcium channel subtypes differ at two types of cholinergic synapse in lumbar sympathetic neurones of guinea-pigs. J. Physiol., 514, 59–69.CrossRefGoogle ScholarPubMed
Ivanov, A. and Purves, D. (1989). Ongoing electrical activity of superior cervical ganglion cells in mammals of different size. J. Comp. Neurol., 284, 398–404.CrossRefGoogle ScholarPubMed
Iversen, S., Iversen, L. and Saper, C. B. (2000). The autonomic nervous system. In The Principles of Neural Science, 4th edn., ed. Kandel, E. R., Schwartz, J. H. and Jessel, T. M.. New York: McGraw-Hill, pp. 960–981.Google Scholar
Izumi, H. (1999). Nervous control of blood flow in the orofacial region. Pharmacol. Ther., 81, 141–161.CrossRefGoogle ScholarPubMed
Izzo, P. N. and Spyer, K. M. (1997). Parasympathetic innervation of the heart. In The Autonomic Nervous System, Vol. 11, Central Nervous Control of Autonomic function, ed. Jordan, D.. Amsterdam: Harwood Academic Publishers, pp. 109–127.Google Scholar
Izzo, P. N., Deuchars, J. and Spyer, K. M. (1993). Localization of cardiac vagal preganglionic motoneurones in the rat: immunocytochemical evidence of synaptic inputs containing 5- hydroxytryptamine. J. Comp. Neurol., 327, 572–583.CrossRefGoogle ScholarPubMed
Jack, J. B. B., Noble, D. and Tsien, R. W. (1975). Electrical Current Flow in Excitable Cells. Oxford: Clarendon Press.Google Scholar
James, W. (1884). What is an emotion?Mind, 9, 188–205.CrossRefGoogle Scholar
James, W. (1994) [1894]. The physical bases of emotion. 1894. Psychol. Rev., 101, 205–210.CrossRefGoogle ScholarPubMed
Jamieson, J., Boyd, H. D. and McLachlan, E. M. (2003). Simulations to derive membrane resistivity in three phenotypes of guinea pig sympathetic postganglionic neuron. J. Neurophysiol., 89, 2430–2440.CrossRefGoogle ScholarPubMed
Jan, L. Y. and Jan, Y. N. (1982). Peptidergic transmission in sympathetic ganglia of the frog. J. Physiol., 327, 219–246.CrossRefGoogle ScholarPubMed
Jan, Y. N., Jan, L. Y. and Kuffler, S. W. (1979). A peptide as a possible transmitter in sympathetic ganglia of the frog. Proc. Natl. Acad. Sci. U.S.A., 76, 1501–1505.CrossRefGoogle ScholarPubMed
Jan, L. Y., Jan, Y. N. and Brownfield, M. S. (1980a). Peptidergic transmitters in synaptic boutons of sympathetic ganglia. Nature, 288, 380–382.CrossRefGoogle Scholar
Jan, Y. N., Jan, L. Y. and Kuffler, S. W. (1980b). Further evidence for peptidergic transmission in sympathetic ganglia. Proc. Natl. Acad. Sci. U.S.A., 77, 5008–5012.CrossRefGoogle Scholar
Jancsó, N. (1960). Role of the nerve terminals in the mechanism of inflammatory reactions. Bull. Millard Fillmore Hosp. (Buffalo, NY), 7, 53–77.Google Scholar
Jancsó, N., Jancsó-Gábor, A. and Szolcsányi, J. (1967). Direct evidence for neurogenic inflammation and its prevention by denervation and by pretreatment with capsaicin. Br. J. Pharmacol., 31, 138–151.Google ScholarPubMed
Jancsó, N., Jancsó-Gábor, A. and Szolcsányi, J. (1968). The role of sensory nerve endings in neurogenic inflammation induced in human skin and in the eye and paw of the rat. Br. J. Pharmacol., 33, 32–41.Google ScholarPubMed
Jänig, W. (1975). Central organization of somatosympathetic reflexes in vasoconstrictor neurones. Brain Res., 87, 305–312.CrossRefGoogle ScholarPubMed
Jänig, W. (1985a). Organization of the lumbar sympathetic outflow to skeletal muscle and skin of the cat hindlimb and tail. Rev. Physiol. Biochem. Pharmacol., 102, 119–213.CrossRefGoogle Scholar
Jänig, W. (1985b). Causalgia and reflex sympathetic dystrophy: in which way is the sympathetic nervous system involved. Trends Neurosci., 8, 471–477.CrossRefGoogle Scholar
Jänig, W. (1986). Spinal cord integration of visceral sensory systems and sympathetic nervous system reflexes. Prog. Brain Res., 67, 255–277.CrossRefGoogle ScholarPubMed
Jänig, W. (1988a). Pre- and postganglionic vasoconstrictor neurons: differentiation, types, and discharge properties. Annu. Rev. Physiol., 50, 525–539.CrossRefGoogle Scholar
Jänig, W. (1988b). Integration of gut function by sympathetic reflexes. Bailliere's Clin. Gastroenterol., 2, 45–62.CrossRefGoogle Scholar
Jänig, W. (1990a). Functions of the sympathetic innervation of the skin. In Central Regulation of Autonomic Function, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 334–348.Google Scholar
Jänig, W. (1990b). The sympathetic nervous system in pain: physiology and pathophysiology. In Pain and the Sympathetic Nervous System, ed. Stanton-Hicks, M.. Boston, Dordrecht, London: Kluwer Academic Publishers, pp. 17–89.Google Scholar
Jänig, W. (1993). Spinal visceral afferents, sympathetic nervous system and referred pain. In New Trends in Referred Pain and Hyperalgesia, Pain Research and Clinical Management, Vol. 7, ed. Vecchiet, L., Albe-Fessard, D., Lindblom, U. and Giamberardino, M. A.. Amsterdam: Elsevier Science Publishers, pp. 83–92.Google Scholar
Jänig, W. (1995a). Ganglionic transmission in vivo. In The Autonomic Nervous System, Vol. 6, Autonomic Ganglia, ed. McLachlan, E. M.. Chur: Harwood Academic Publishers, pp. 349–395.Google Scholar
Jänig, W. (1995b). The sympathetic nervous system in pain. Eur. J. Anaesthesiol., 12 (Suppl. 10), 53–60.Google Scholar
Jänig, W. (1996a). Spinal cord reflex organization of sympathetic systems. Prog. Brain Res., 107, 43–77.CrossRefGoogle Scholar
Jänig, W. (1996b). Regulation of the lower urinary tract. In Comprehensive Human Physiology, Vol. 2, ed. Greger, R. and Windhorst, U.. Berlin, Heidelberg: Springer-Verlag, pp. 1611–1624.CrossRefGoogle Scholar
Jänig, W. (1996c). Behavioral and neurovegetative components of reproductive functions. In Comprehensive Human Physiology, Vol. 2, ed. Greger, R. and Windhorst, U.. Berlin, Heidelberg: Springer-Verlag, pp. 2253–2263.CrossRefGoogle Scholar
Jänig, W. (1996d). Neurobiology of visceral afferent neurons: neuroanatomy, functions, organ regulations and sensations. Biol. Psychol., 42, 29–51.CrossRefGoogle Scholar
Jänig, W. (2002). Pain in the sympathetic nervous system: pathophysiological mechanisms. In Autonomic Failure, 4th edn., ed. Bannister, R. and Mathias, C. J.. New York, Oxford: Oxford University Press, pp. 99–108.Google Scholar
Jänig, W. (2005a). Vegetatives Nervensystem. In Physiologie des Menschen, 29th edn., ed. Schmidt, R. F., Lang, F. and Thews, G.. Heidelberg Berlin: Springer Medizin Verlag, pp. 425–458.CrossRefGoogle Scholar
Jänig, W. (2005b). Vagal afferents and visceral pain. In Advances in Vagal Afferent Neurobiology, ed. Undem, B. and Weinreich, D.. Boca Raton: CRC Press, pp. 461–489.CrossRefGoogle Scholar
Jänig, W. and Baron, R. (2001). The role of the sympathetic nervous system in neuropathic pain: clinical observations and animal models. In Neuropathic Pain: Pathophysiological and Treatment, ed. Hansson, P. T., Fields, H. L., Hill, R. G. and Marchettini, P.. Seattle: IASP Press, pp. 125–149.Google Scholar
Jänig, W. and Baron, R. (2002). Complex regional pain syndrome is a disease of the central nervous system. Clin. Auton. Res., 12, 150–164.CrossRefGoogle ScholarPubMed
Jänig, W. and Baron, R. (2003). Complex regional pain syndrome: mystery explained? Lancet Neurol., 2, 687–697.CrossRefGoogle ScholarPubMed
Jänig, W. and Häbler, H. J. (1995). Visceral-autonomic integration. In Visceral Pain; Progress in Pain Research and Management, Vol. 5, ed. Gebhart, G. F.. Seattle: IASP Press, pp. 311–348.Google Scholar
Jänig, W. and Häbler, H. J. (1999). Organisation of the autonomic nervous system: structure and function. In Handbook of Clinical Neurology, Vol. 74, The Autonomic Nervous System, part I: Normal Functions, ed. Appenzeller, O.. Amsterdam: Elsevier, pp. 1–52.Google Scholar
Jänig, W. and Häbler, H. J. (2000a). Specificity in the organization of the autonomic nervous system: a basis for precise neural regulation of homeostatic and protective body functions. Prog. Brain Res., 122, 351–367.CrossRefGoogle Scholar
Jänig, W. and Häbler, H. J. (2000b). Sympathetic nervous system: contribution to chronic pain. Prog. Brain Res., 129, 451–468.Google Scholar
Jänig, W. and Häbler, H. J. (2002). Physiologie und pathophysiologie viszeraler Schmerzen [Physiology and pathophysiology of visceral pain]. Schmerz., 16, 429–446.CrossRefGoogle Scholar
Jänig, W. and Häbler, H. J. (2003). Neurophysiological analysis of target-related sympathetic pathways – from animal to human: similarities and differences. Acta Physiol. Scand., 177, 255–274.CrossRefGoogle ScholarPubMed
Jänig, W. and Koltzenburg, M. (1990). On the function of spinal primary afferent fibres supplying colon and urinary bladder. J. Auton. Nerv. Syst., 30 Suppl, S89–S96.CrossRefGoogle ScholarPubMed
Jänig, W. and Koltzenburg, M. (1991a). What is the interaction between the sympathetic terminal and the primary afferent fiber? In Towards a New Pharmacotherapy of Pain, ed. Basbaum, A. I. and Besson, J.-M.. Chichester: Dahlem Workshop Reports John Wiley & Sons, pp. 331–352.Google Scholar
Jänig, W. and Koltzenburg, M. (1991b). Plasticity of sympathetic reflex organization following cross-union of inappropriate nerves in the adult cat. J. Physiol., 436, 309–323.CrossRefGoogle Scholar
Jänig, W. and Koltzenburg, M. (1991c). Receptive properties of sacral primary afferent neurons supplying the colon. J. Neurophysiol., 65, 1067–1077.CrossRefGoogle Scholar
Jänig, W. and Koltzenburg, M. (1993). Pain arising from the urogenital tract. In The Autonomic Nervous System, Vol. 3., Nervous Control of the Urogenital Tract, ed. Maggi, C. A.. Chur, Switzerland: Harwood Academic Publisher, pp. 523–576.Google Scholar
Jänig, W. and Kümmel, H. (1977). Functional discrimination of postganglionic neurones to the cat's hindpaw with respect to the skin potentials recorded from the hairless skin. Pflügers Arch., 371, 217–225.CrossRefGoogle ScholarPubMed
Jänig, W. and Kümmel, H. (1981). Organization of the sympathetic innervation supplying the hairless skin of the cat's paw. J. Auton. Nerv. Syst., 3, 215–230.CrossRefGoogle ScholarPubMed
Jänig, W. and Levine, J. D. (2005). Autonomic-endocrine-immune responses in acute and chronic pain. In Wall and Melzack's Textbook of Pain, 5th edn., ed. McMahon, S. B. and Koltzenburg, M.. Edinburgh: Elsevier Churchill Livingstone, pp. 205–218.Google Scholar
Jänig, W. and Lisney, S. J. (1989). Small diameter myelinated afferents produce vasodilation but not plasma extravasation in rat skin. J. Physiol., 415, 477–486.CrossRefGoogle Scholar
Jänig, W. and McLachlan, E. M. (1986a). The sympathetic and sensory components of the caudal lumbar sympathetic trunk in the cat. J. Comp. Neurol., 245, 62–73.CrossRefGoogle Scholar
Jänig, W. and McLachlan, E. M. (1986b). Identification of distinct topographical distributions of lumbar sympathetic and sensory neurons projecting to end organs with different functions in the cat. J. Comp. Neurol., 246, 104–112.CrossRefGoogle Scholar
Jänig, W. and McLachlan, E. M. (1987). Organization of lumbar spinal outflow to distal colon and pelvic organs. Physiol. Rev., 67, 1332–1404.CrossRefGoogle ScholarPubMed
Jänig, W. and McLachlan, E. M. (1992a). Characteristics of function-specific pathways in the sympathetic nervous system. Trends Neurosci., 15, 475–481.CrossRefGoogle Scholar
Jänig, W. and McLachlan, E. M. (1992b). Specialized functional pathways are the building blocks of the autonomic nervous system. J. Auton. Nerv. Syst., 41, 3–13.CrossRefGoogle Scholar
Jänig, W. and McLachlan, E. M. (1994). The role of modifications in noradrenergic peripheral pathway after nerve lesions in the generation of pain. In Pharmacological Approaches to the Treatment of Pain: New Concepts and Critical Issues, Progress in Pain Research and Management, Vol. 1, ed. Fields, H. L. and Liebeskind, J. C.. Seattle: IASP Press, pp. 101–128.Google Scholar
Jänig, W. and McLachlan, E. M. (2002). Neurobiology of the autonomic nervous system. In Autonomic Failure, 4th edn., ed. Mathias, C. J. and Bannister, R.. New York, Oxford: Oxford University Press, pp. 3–15.Google ScholarPubMed
Jänig, W. and Morrison, J. F. B. (1986). Functional properties of spinal visceral afferents supplying abdominal and pelvic organs, with special emphasis on visceral nociception. Prog. Brain Res., 67, 87–114.CrossRefGoogle ScholarPubMed
Jänig, W. and Räth, B. (1977). Electrodermal reflexes in the cat's paws elicited by natural stimulation of skin. Pflügers Arch., 369, 27–32.CrossRefGoogle Scholar
Jänig, W. and Räth, B. (1980). Effects of anaesthetics on reflexes elicited in the sudomotor system by stimulation of Pacinian corpuscles and of cutaneous nociceptors. J. Auton. Nerv. Syst., 2, 1–14.CrossRefGoogle ScholarPubMed
Jänig, W. and Spilok, N. (1978). Functional organization of the sympathetic innervation supplying the hairless skin of the hindpaws in chronic spinal cats. Pflügers Arch., 377, 25–31.CrossRefGoogle ScholarPubMed
Jänig, W. and Stanton-Hicks, M. (eds.) (1996). Reflex Sympathetic Dystrophy – a Reappraisal. Seattle: IASP Press.Google Scholar
Jänig, W. and Szulczyk, P. (1980). Functional properties of lumbar preganglionic neurones. Brain Res., 186, 115–131.CrossRefGoogle ScholarPubMed
Jänig, W. and Szulczyk, P. (1981). The organization of lumbar preganglionic neurons. J. Auton. Nerv. Syst., 3, 177–191.CrossRefGoogle ScholarPubMed
Jänig, W., Krauspe, R. and Wiedersatz, G. (1982). Transmission of impulses from pre- to postganglionic vasoconstrictor and sudomotor neurons. J. Auton. Nerv. Syst., 6, 95–106.CrossRefGoogle ScholarPubMed
Jänig, W., Krauspe, R. and Wiedersatz, G. (1983). Reflex activation of postganglionic vasoconstrictor neurones supplying skeletal muscle by stimulation of arterial chemoreceptors via non-nicotinic synaptic mechanisms in sympathetic ganglia. Pflügers Arch., 396, 95–100.CrossRefGoogle ScholarPubMed
Jänig, W., Krauspe, R. and Wiedersatz, G. (1984). Activation of postganglionic neurones via non-nicotinic synaptic mechanisms by stimulation of thin preganglionic axons. Pflügers Arch., 401, 318–320.CrossRefGoogle ScholarPubMed
Jänig, W., Schmidt, M., Schnitzler, A. and Wesselmann, U. (1991). Differentiation of sympathetic neurones projecting in the hypogastric nerves in terms of their discharge patterns in cats. J. Physiol., 437, 157–179.CrossRefGoogle ScholarPubMed
Jänig, W., Levine, J. D. and Michaelis, M. (1996). Interactions of sympathetic and primary afferent neurons following nerve injury and tissue trauma. Prog. Brain Res., 112, 161–184.CrossRefGoogle Scholar
Jänig, W., Khasar, S. G., Levine, J. D. and Miao, F. J. P. (2000). The role of vagal visceral afferents in the control of nociception. Prog. Brain Res., 122, 273–287.CrossRefGoogle ScholarPubMed
Jänig, W., Chapman, C. R. and Green, P. G. (2006). Pain and body protection: sensory, autonomic, neuroendocrine and behavioral mechanisms in the control of inflammation and hyperalgesia. In Proceeding of the 11th World Congress on Pain, ed. Flor, H., Kalso, E. and Dostrovsky, J. O.. Seattle: IASP Press, in press.Google Scholar
Jankowska, E. (1992). Interneuronal relay in spinal pathways from proprioceptors. Prog. Neurobiol., 38, 335–378.CrossRefGoogle ScholarPubMed
Jankowska, E. (2001). Spinal interneuronal systems: identification, multifunctional character and reconfigurations in mammals. J. Physiol., 533, 31–40.CrossRefGoogle ScholarPubMed
Jankowska, E. and Lundberg, A. (1981). Interneurones in the spinal cord. Trends Neurosci., 4, 230–233.CrossRefGoogle Scholar
Jansen, A. S. and Loewy, A. D. (1997). Neurons lying in the white matter of the upper cervical spinal cord project to the intermediolateral cell column. Neuroscience, 77, 889–898.CrossRefGoogle ScholarPubMed
Jansen, A. S., Horst, G. J., Mettenleiter, T. C. and Loewy, A. D. (1992). central nervous system cell groups projecting to the submandibular parasympathetic preganglionic neurons in the rat: a retrograde transneuronal viral cell body labeling study. Brain Res., 572, 253–260.CrossRefGoogle ScholarPubMed
Jansen, A. S., Farwell, D. G. and Loewy, A. D. (1993). Specificity of pseudorabies virus as a retrograde marker of sympathetic preganglionic neurons: implications for transneuronal labeling studies. Brain Res., 617, 103–112.CrossRefGoogle ScholarPubMed
Jansen, A. S., Nguyen, X. V., Karpitskiy, V., Mettenleiter, T. C. and Loewy, A. D. (1995a). Central command neurons of the sympathetic nervous system: basis of the fight-or-flight response. Science, 270, 644–646.CrossRefGoogle Scholar
Jansen, A. S., Wessendorf, M. W. and Loewy, A. D. (1995b). Transneuronal labeling of central nervous system neuropeptide and monoamine neurons after pseudorabies virus injections into the stellate ganglion. Brain Res., 683, 1–24.CrossRefGoogle Scholar
Jeske, I., Morrison, S. F., Cravo, S. L. and Reis, D. J. (1993). Identification of baroreceptor reflex interneurons in the caudal ventrolateral medulla. Am. J. Physiol., 264, R169–R178.Google ScholarPubMed
Jewett, D. L. (1964). Activity of single efferent fibres in the cervical vagus nerve of the dog, with special reference to possible cardio-inhibitory fibres. J. Physiol., 175, 321–357.CrossRefGoogle ScholarPubMed
Jichia, D. and Frank, J. L. (2000). Spinal cord disease trauma and autonomic nervous system dysfunction. In Handbook of Clinical Neurology, Vol. 75, The Autonomic Nervous System part II: Dysfunctions, ed. Appenzeller, O., Amsterdam: Elsevier Science, pp. 567–587.Google Scholar
Jobling, P. and McLachlan, E. M. (1992). An electrophysiological study of responses evoked in isolated segments of rat tail artery during growth and maturation. J. Physiol., 454, 83–105.CrossRefGoogle ScholarPubMed
Jobling, P., McLachlan, E. M., Jänig, W. and Anderson, C. R. (1992). Electrophysiological responses in the rat tail artery during reinnervation following lesions of the sympathetic supply. J. Physiol., 454, 107–128.CrossRefGoogle ScholarPubMed
Jobling, P., Gibbins, I. L. and Morris, J. L. (2003). Functional organization of vasodilator neurons in pelvic ganglia of female guinea pigs: comparison with uterine motor neurons. J. Comp. Neurol., 459, 223–241.CrossRefGoogle ScholarPubMed
Johnson, C. D. and Gilbey, M. P. (1996). On the dominant rhythm in the discharges of single postganglionic sympathetic neurones innervating the rat tail artery. J. Physiol., 497, 241–259.CrossRefGoogle ScholarPubMed
Johnson, D. A. and Purves, D. (1981). Post-natal reduction of neural unit size in the rabbit ciliary ganglion. J. Physiol., 318, 143–159.CrossRefGoogle ScholarPubMed
Johnson, D. A. and Purves, D. (1983). Tonic and reflex synaptic activity recorded in ciliary ganglion cells of anaesthetized rabbits. J. Physiol., 339, 599–613.CrossRefGoogle ScholarPubMed
Jones, E. G. (2002). A pain in the thalamus. J. Pain, 3, 102–104.CrossRefGoogle ScholarPubMed
Jones, E. G. (2006). The Thalamus, 2nd edn., 2 volumes. Cambridge: Cambridge University Press.Google Scholar
Jones, J. F., Wang, Y. and Jordan, D. (1998). Activity of C fibre cardiac vagal efferents in anaesthetized cats and rats. J. Physiol., 507, 869–880.CrossRefGoogle ScholarPubMed
Jordan, D. (1997a). Central nervous control of the airways. In The Autonomic Nervous System, Vol. 11, Central Nervous Control of Autonomic Function, ed. Jordan, D.. Amsterdam: Harwood Academic Publishers, pp. 63–107.Google Scholar
Jordan, D. (ed.) (1997b). The Autonomic Nervous System, Vol. 11, Central Nervous Control of Autonomic Function, Amsterdam: Harwood Academic Press.
Joyner, M. J. and Dietz, N. M. (2003). Sympathetic vasodilation in human muscle. Acta Physiol. Scand., 177, 329–336.CrossRefGoogle ScholarPubMed
Joyner, M. J. and Halliwill, J. R. (2000). Sympathetic vasodilatation in human limbs. J. Physiol., 526, 471–480.Google ScholarPubMed
Julé, Y. and Szurszewski, J. H. (1983). Electrophysiology of neurones of the inferior mesenteric ganglion of the cat. J. Physiol., 344, 277–292.CrossRefGoogle ScholarPubMed
Juler, G. L. and Eltorai, I. M. (1985). The acute abdomen in spinal cord injury patients. Paraplegia, 23, 118–123.Google ScholarPubMed
Kalia, M. and Richter, D. (1985a). Morphology of physiologically identified slowly adapting lung stretch receptor afferents stained with intra-axonal horseradish peroxidase in the nucleus of the tractus solitarius of the cat. I. A light microscopic analysis. J. Comp. Neurol., 241, 503–520.CrossRefGoogle Scholar
Kalia, M. and Richter, D. (1985b). Morphology of physiologically identified slowly adapting lung stretch receptor afferents stained with intra-axonal horseradish peroxidase in the nucleus of the tractus solitarius of the cat. II. An ultrastructural analysis. J. Comp. Neurol., 241, 521–535.CrossRefGoogle Scholar
Kalia, M. and Richter, D. (1988a). Rapidly adapting pulmonary receptor afferents: I. Arborization in the nucleus of the tractus solitarius. J. Comp. Neurol., 274, 560–573.CrossRefGoogle Scholar
Kalia, M. and Richter, D. (1988b). Rapidly adapting pulmonary receptor afferents: II. Fine structure and synaptic organization of central terminal processes in the nucleus of the tractus solitarius. J. Comp. Neurol., 274, 574–594.CrossRefGoogle Scholar
Kanosue, K., Hosono, T., Zhang, Y. H. and Chen, X. M. (1998). Neuronal networks controlling thermoregulatory effectors. Prog. Brain Res., 115, 49–62.CrossRefGoogle Scholar
Karczmar, K., Koketsu, K. and Nishi, S. (eds.) (1986). Autonomic and Enteric Ganglia. New York: Plenum Press.CrossRefGoogle Scholar
Karlsson, A. K. (1999). Autonomic dysreflexia. Spinal Cord, 37, 383–391.CrossRefGoogle ScholarPubMed
Katayama, Y. and Nishi, S. (1986). Peptidergic transmission. In Autonomic and Enteric Ganglia, ed. Karczmar, K., Koketsu, K. and Nishi, S.. New York, London: Plenum Press, pp. 181–200.CrossRefGoogle Scholar
Katona, P. G., Poitras, J. W., Barnett, G. O. and Terry, B. S. (1970). Cardiac vagal efferent activity and heart period in the carotid sinus reflex. Am. J. Physiol., 218, 1030–1037.Google ScholarPubMed
Katter, J. T., Dado, R. J., Kostarczyk, E. and Giesler, G. J. Jr. (1996). Spinothalamic and spinohypothalamic tract neurons in the sacral spinal cord of rats. II. Responses to cutaneous and visceral stimuli. J. Neurophysiol., 75, 2606–2628.CrossRefGoogle ScholarPubMed
Kazuyuki, K., Hosono, T., Zhang, Y. H. and Chen, X. M. (1998). Neuronal networks controlling thermoregulatory effectors. Prog. Brain Res., 115, 49–62.Google ScholarPubMed
Keast, J. R. (1995a). Pelvic ganglia. In The Autonomic Nervous System, Vol. 6, Autonomic Ganglia, ed. McLachlan, E. M.. London: Harwood Academic Publishers GmbH, pp. 445–479.Google Scholar
Keast, J. R. (1995b). Visualization and immunohistochemical characterization of sympathetic and parasympathetic neurons in the male rat major pelvic ganglion. Neuroscience, 66, 655–662.CrossRefGoogle Scholar
Keast, J. R. (1999). Unusual autonomic ganglia: connections, chemistry, and plasticity of pelvic ganglia. Int. Rev. Cytol., 193, 1–69.CrossRefGoogle ScholarPubMed
Keast, J. R., McLachlan, E. M. and Meckler, R. L. (1993). Relation between electrophysiological class and neuropeptide content of guinea pig sympathetic prevertebral neurons. J. Neurophysiol., 69, 384–394.CrossRefGoogle ScholarPubMed
Keast, J. R., Luckensmeyer, G. B. and Schemann, M. (1995). All pelvic neurons in male rats contain immunoreactivity for the synthetic enzymes of either noradrenaline or acetylcholine. Neurosci. Lett., 196, 209–212.CrossRefGoogle ScholarPubMed
Keay, K. A. and Bandler, R. (2001). Parallel circuits mediating distinct emotional coping reactions to different types of stress. Neurosci. Biobehav. Rev., 25, 669–678.CrossRefGoogle ScholarPubMed
Keay, K. A. and Bandler, R. (2004). Periaqueductal gray. In The Rat Nervous System, 3rd edn., ed. Paxinos, G.. San Diego: Academic Press, pp. 243–257.Google Scholar
Keay, K. A., Clement, C. I., Owler, B., Depaulis, A. and Bandler, R. (1994). Convergence of deep somatic and visceral nociceptive information onto a discrete ventrolateral midbrain periaqueductal gray region. Neuroscience, 61, 727–732.CrossRefGoogle ScholarPubMed
Keay, K. A., Feil, K., Gordon, B. D., Herbert, H. and Bandler, R. (1997). Spinal afferents to functionally distinct periaqueductal gray columns in the rat: an anterograde and retrograde tracing study. J. Comp. Neurol., 385, 207–229.3.0.CO;2-5>CrossRefGoogle ScholarPubMed
Keay, K. A., Clement, C. I., Matar, W. M.et al. (2002). Noxious activation of spinal or vagal afferents evokes distinct patterns of fos-like immunoreactivity in the ventrolateral periaqueductal gray of unanaesthetised rats. Brain Res., 948, 122–130.CrossRefGoogle ScholarPubMed
Keef, K. D. and Kreulen, D. L. (1986). Venous mechanoreceptor input to neurones in the inferior mesenteric ganglion of the guinea-pig. J. Physiol., 377, 49–59.CrossRefGoogle ScholarPubMed
Kellogg, D. L. Jr., Johnson, J. M. and Kosiba, W. A. (1989). Selective abolition of adrenergic vasoconstrictor responses in skin by local iontophoresis of bretylium. Am. J. Physiol., 257, H1599–H1606.Google ScholarPubMed
Kellogg, D. L. Jr., Pergola, P. E., Piest, K. L.et al. (1995). Cutaneous active vasodilation in humans is mediated by cholinergic nerve cotransmission. Circ. Res., 77, 1222–1228.CrossRefGoogle ScholarPubMed
Kesler, B. S., Mazzone, S. B. and Canning, B. J. (2002). Nitric oxide-dependent modulation of smooth-muscle tone by airway parasympathetic nerves. Am. J. Respir. Crit. Care, 165, 481–488.CrossRefGoogle ScholarPubMed
Khasar, S. G., Green, P. G. and Levine, J. D. (1993). Comparison of intradermal and subcutaneous hyperalgesic effects of inflammatory mediators in the rat. Neurosci. Lett., 153, 215–218.CrossRefGoogle ScholarPubMed
Khasar, S. G., Miao, F. J. P. and Levine, J. D. (1995). Inflammation modulates the contribution of receptor-subtypes to bradykinin-induced hyperalgesia in the rat. Neuroscience, 69, 685–690.CrossRefGoogle ScholarPubMed
Khasar, S. G., Miao, F. J. P., Jänig, W. and Levine, J. D. (1998a). Modulation of bradykinin-induced mechanical hyperalgesia in the rat by activity in abdominal vagal afferents. Eur. J. Neurosci., 10, 435–444.CrossRefGoogle Scholar
Khasar, S. G., Miao, F. J. P., Jänig, W. and Levine, J. D. (1998b). Vagotomy-induced enhancement of mechanical hyperalgesia in the rat is sympathoadrenal-mediated. J. Neurosci., 18, 3043–3049.CrossRefGoogle Scholar
Khasar, S. G., Green, P. G., Miao, F. J. P. and Levine, J. D. (2003). Vagal modulation of nociception is mediated by adrenomedullary epinephrine in the rat. Eur. J. Neurosci., 17, 909–915.CrossRefGoogle ScholarPubMed
Kim, M., Chiego, D. J. Jr. and Bradley, R. M. (2004). Morphology of parasympathetic neurons innervating rat lingual salivary glands. Auton. Neurosci., 111, 27–36.CrossRefGoogle ScholarPubMed
Kirchheim, H. R., Just, A. and Ehmke, H. (1998). Physiology and pathophysiology of baroreceptor function and neuro-hormonal abnormalities in heart failure. Basic Res. Cardiol., 93, Suppl. 1, 1–22.CrossRefGoogle ScholarPubMed
Kirkup, A. J., Brunsden, A. M. and Grundy, D. (2001). Receptors and transmission in the brain-gut axis: potential for novel therapies. I. Receptors on visceral afferents. Am. J. Physiol. Gastrointest. Liver Physiol., 280, G787–G794.CrossRefGoogle ScholarPubMed
Klemm, M. F. (1995). Neuromuscular junctions made by nerve fibres supplying the longitudinal muscle of the guinea-pig ileum. J. Auton. Nerv. Syst., 55, 155–164.CrossRefGoogle ScholarPubMed
Klemm, M., Hirst, G. D. and Campbell, G. (1992). Structure of autonomic neuromuscular junctions in the sinus venosus of the toad. J. Auton. Nerv. Syst., 39, 139–150.CrossRefGoogle ScholarPubMed
Klemm, M. F., Helden, D. F. and Luff, S. E. (1993). Ultrastructural analysis of sympathetic neuromuscular junctions on mesenteric veins of the guinea pig. J. Comp. Neurol., 334, 159–167.CrossRefGoogle ScholarPubMed
Koepchen, H. P. (1962). Die Blutdruckrhythmik [The Rhythm of Blood Pressure]. Darmstadt: Dr. Dietrich Steinkopff.Google Scholar
Koepchen, H. P. (1983). Respiratory and cardiovascular “centres”: functional entirety or separate structures. In Central Environment and the Control Systems of Breathing and Circulation, ed. Schläfke, M. E., Koepchen, H. P. and See, W. R.. Berlin, Heidelberg, New York: Springer, pp. 221–237.CrossRefGoogle Scholar
Koepchen, H. P. and Thurau, K. (1959). Über die Entstehungsbedingungen der atemsynchronen Schwankungen des Vagustonus (respiratorische Arrhythmie) [On the origin of the respiratory synchronous changes of the vagus tone (respiratory arrhythmia)]. Pflügers Arch., 259, 10–30.CrossRefGoogle Scholar
Koepchen, H. P., Seller, H., Polster, J. and Langhorst, P. (1968). Über die Fein-Vasomotorik der Muskelstrombahn und ihre Beziehung zur Ateminnervation [Spontaneous vasomotor changes in the muscle and their relation to the respiratory rhythm]. Pflügers Arch., 302, 285–299.CrossRefGoogle Scholar
Koepchen, H. P., Klüssendorf, D. and Sommer, D. (1981). Neurophysiological background of central neural cardiovascular-respiratory coordination: basic remarks and experimental approach. J. Auton. Nerv. Syst., 3, 335–368.CrossRefGoogle ScholarPubMed
Koepchen, H. P., Abel, H.-H. and Klüssendorf, D. (1987). Brain stem generation of specific and non-specific rhythms. In Organization of the Automatic Nervous System: Central and Peripheral Mechanisms, ed. Ciriello, J., Calaresu, F. R., Renaud, L. P. and Polosa, C.. New York: Alan R. Liss., pp. 179–188.Google Scholar
Koh, S. D., Ward, S. M., Ordog, T., Sanders, K. M. and Horowitz, B. (2003). Conductances responsible for slow wave generation and propagation in interstitial cells of Cajal. Curr. Opin. Pharmacol., 3, 579–582.CrossRefGoogle ScholarPubMed
Koltzenburg, M., Häbler, H. J. and Jänig, W. (1995). Functional reinnervation of the vasculature of the adult cat paw pad by axons originally innervating vessels in hairy skin. Neuroscience, 67, 245–252.CrossRefGoogle ScholarPubMed
Kopin, I. J. (1989). Plasma levels of catecholamines and dopamine-beta-hydroxylase. In Handbook of Experimental Pharmacology, Vol. 90/II, Catecholamines II, ed. Trendelenburg, U. and Weiner, N.. Berlin: Springer-Verlag, pp. 211–275.Google Scholar
Kopp, U. C. and DiBona, G. F. (2000). The neural control of renal function. In The Kidney: Physiology and Pathophysiology, 3rd edn., ed. Seldin, G. and Giebisch, G.. New York: Raven Press, pp. 981–1006.Google Scholar
Korner, P. I. (1979). Central nervous control of autonomic cardiovascular function. In Handbook of Physiology, The Cardiovascular System, Vol. I, The Heart. ed. Barne, R. M.. Bethesda: American Physiological Society, pp. 691–739.Google Scholar
Koshiya, N. and Guyenet, P. G. (1996). Tonic sympathetic chemoreflex after blockade of respiratory rhythmogenesis in the rat. J. Physiol., 491, 859–869.CrossRefGoogle ScholarPubMed
Koshiya, N., Huangfu, D. and Guyenet, P. G. (1993). Ventrolateral medulla and sympathetic chemoreflex in the rat. Brain Res., 609, 174–184.CrossRefGoogle ScholarPubMed
Kostarczyk, E., Zhang, X. and Giesler, G. J. Jr. (1997). Spinohypothalamic tract neurons in the cervical enlargement of rats: locations of antidromically identified ascending axons and their collateral branches in the contralateral brain. J. Neurophysiol., 77, 435–451.CrossRefGoogle ScholarPubMed
Koushanpour, E. (1991). Baroreceptor discharge behavior and resetting. In Baroreceptor Reflexes, ed. Persson, P. B. and Kirchheim, H. R.. Berlin, Heidelberg: Springer-Verlag, pp. 9–44.CrossRefGoogle Scholar
Krassioukov, A. V., Johns, D. G. and Schramm, L. P. (2002). Sensitivity of sympathetically correlated spinal interneurons, renal sympathetic nerve activity, and arterial pressure to somatic and visceral stimuli after chronic spinal injury. J. Neurotrauma, 19, 1521–1529.CrossRefGoogle ScholarPubMed
Kreis, M. E., Jiang, W., Kirkup, A. J. and Grundy, D. (2002). Cosensitivity of vagal mucosal afferents to histamine and 5-hydroxytryptamine (serotonin) in the rat jejunum. Am. J. Physiol. Gastrointest. Liver Physiol., 283, G612–G617.CrossRefGoogle ScholarPubMed
Krenz, N. R. and Weaver, L. C. (2000). Nerve growth factor in glia and inflammatory cells of the injured rat spinal cord. J. Neurochem., 74, 730–739.CrossRefGoogle ScholarPubMed
Krenz, N. R., Meakin, S. O., Krassioukov, A. V. and Weaver, L. C. (1999). Neutralizing intraspinal nerve growth factor blocks autonomic dysreflexia caused by spinal cord injury. J. Neurosci., 19, 7405–7414.CrossRefGoogle ScholarPubMed
Kress, M., Koltzenburg, M., Reeh, P. W. and Handwerker, H. O. (1992). Responsiveness and functional attributes of electrically localized terminals of cutaneous C-fibers in vivo and in vitro. J. Neurophysiol., 68, 581–595.CrossRefGoogle ScholarPubMed
Kreulen, D. L. and Peters, S. (1986). Non-cholinergic transmission in a sympathetic ganglion of the guinea-pig elicited by colon distension. J. Physiol., 374, 315–334.CrossRefGoogle Scholar
Kreulen, D. L. and Szurszewski, J. H. (1979a). Nerve pathways in celiac plexus of the guinea pig. Am. J. Physiol., 237, E90–E97.Google Scholar
Kreulen, D. L. and Szurszewski, J. H. (1979b). Reflex pathways in the abdominal prevertebral ganglia: evidence for a colo-colonic inhibitory reflex. J. Physiol., 295, 21–32.CrossRefGoogle Scholar
Krier, J. and Hartman, D. A. (1984). Electrical properties and synaptic connections to neurons in parasympathetic colonic ganglia of the cat. Am. J. Physiol., 247, G52–G61.Google ScholarPubMed
Krier, J. and Szurszewski, J. H. (1982). Effect of substance P on colonic mechanoreceptors, motility, and sympathetic neurons. Am. J. Physiol., 243, G259–G267.Google ScholarPubMed
Krier, J., Schmalz, P. F. and Szurszewski, J. H. (1982). Central innervation of neurones in the inferior mesenteric ganglion and of the large intestine of the cat. J. Physiol., 332, 125–138.CrossRefGoogle ScholarPubMed
Kuhn, R. A. (1950). Functional capacity of the isolated human spinal cord. Brain, 73, 1–51.CrossRefGoogle ScholarPubMed
Kumazawa, T. (1986). Sensory innervation of reproductive organs. Prog. Brain Res., 67, 115–131.CrossRefGoogle ScholarPubMed
Kumazawa, T. (1990). Functions of the nociceptive primary neurons. Jpn. J. Physiol., 40, 1–14.CrossRefGoogle ScholarPubMed
Kumazawa, T., Mizumura, K. and Sato, J. (1987). Response properties of polymodal receptors studied using in vitro testis superior spermatic nerve preparations of dogs. J. Neurophysiol., 57, 702–711.CrossRefGoogle ScholarPubMed
Kümmel, H. (1983). Activity in sympathetic neurons supplying skin and skeletal muscle in spinal cats. J. Auton. Nerv. Syst., 7, 319–327.CrossRefGoogle ScholarPubMed
Kunitake, T. and Kannan, H. (2000). Discharge pattern of renal sympathetic nerve activity in the conscious rat: spectral analysis of integrated activity. J. Neurophysiol., 84, 2859–2867.CrossRefGoogle ScholarPubMed
Kuntz, A. (1940). The structural organization of the inferior mesenteric ganglia. J. Comp. Neurol., 72, 371–382.CrossRefGoogle Scholar
Kuntz, A. (1954). The Autonomic Nervous System. Philadelphia: Lea & Febinger.Google Scholar
Kuntz, A. and Saccomanno, G. (1944). Reflex inhibition of intestinal motility mediated through decentralized prevertebral ganglia. J. Neurophysiol., 7, 163–171.CrossRefGoogle Scholar
Kunze, D. L. (1972). Reflex discharge patterns of cardiac vagal efferent fibres. J. Physiol., 222, 1–15.CrossRefGoogle ScholarPubMed
Kunze, W. A. and Furness, J. B. (1999). The enteric nervous system and regulation of intestinal motility. Annu. Rev. Physiol., 61, 117–142.CrossRefGoogle ScholarPubMed
Kuo, D. C., Hisamitsu, T. and Groat, W. C. (1984). A sympathetic projection from sacral paravertebral ganglia to the pelvic nerve and to postganglionic nerves on the surface of the urinary bladder and large intestine of the cat. J. Comp. Neurol., 226, 76–86.CrossRefGoogle ScholarPubMed
LaBar, K. S. and LeDoux, J. E. (2001). Coping with danger: the neural basis of defensive behavior and fearful feelings. In Handbook of Physiology. Section 7: The Endocrine System, Vol. IV, Coping with the Environment: Neural and Neuroendocrine Mechanisms, ed. McEwen, B. S.. Oxford, New York: Oxford University Press, pp. 139–154.Google Scholar
Lal, S., Kirkup, A. J., Brunsden, A. M., Thompson, D. G. and Grundy, D. (2001). Vagal afferent responses to fatty acids of different chain length in the rat. Am. J. Physiol. Gastrointest. Liver Physiol., 281, G907–G915.CrossRefGoogle ScholarPubMed
Lamb, K., Kang, Y. M., Gebhart, G. F. and Bielefeldt, K. (2003). Gastric inflammation triggers hypersensitivity to acid in awake rats. Gastroenterology, 125, 1410–1418.CrossRefGoogle ScholarPubMed
Lammers, W. J. (2000). Propagation of individual spikes as “patches” of activation in isolated feline duodenum. Am. J. Physiol. Gastrointest. Liver Physiol., 278, G297–G307.CrossRefGoogle Scholar
LaMotte, R. H., Lundberg, L. E. R. and Torebjörk, H. E. (1992). Pain, hyperalgesia and activity in nociceptive C units in humans after intradermal injection of capsaicin. J. Physiol., 448, 749–764.CrossRefGoogle ScholarPubMed
Lange, C. S. (1920) [1887]. Über Gemüthsbewegungen [translated into English]. In The Emotions, ed. James, W. and Lange, C. G.. Baltimore: Williams and Wilkins.Google Scholar
Langley, J. N. (1891). On the course and connections of the secretory fibres supplying the sweat glands of the feet of the cat. J. Physiol., 12, 347–374.CrossRefGoogle ScholarPubMed
Langley, J. N. (1892). On the origin from the spinal cord of the cervical and upper thoracic sympathetic fibres, with some observations on white and grey rami communicantes. Phil. Trans. R. Soc. Lond. B, 183, 85–124.CrossRefGoogle Scholar
Langley, J. N. (1894a). The arrangement of the sympathetic nervous system, based chiefly on observation upon pilo-motor nerves. J. Physiol., 15, 176–244.CrossRefGoogle Scholar
Langley, J. N. (1894b). Further observations on the secretory and vaso-motor fibres of the foot of the cat, with notes on other sympathetic nerve fibres. J. Physiol., 17, 296–314.CrossRefGoogle Scholar
Langley, J. N. (1895). Note on the regeneration of prae-ganglionic fibres of the sympathetic. J. Physiol., 18, 80–84.CrossRefGoogle ScholarPubMed
Langley, J. N. (1897). On the regeneration of preganglionic and of postganglionic visceral nerve fibres. J. Physiol., 22, 215–230.CrossRefGoogle Scholar
Langley, J. N. (1900). The sympathetic and other related systems of nerves. In Textbook of Physiology, ed. Schäfer, E. A.. Edinburgh, London: Young J. Pentland, pp. 616–696.Google Scholar
Langley, J. N. (1903a). Das sympathische und verwandte nervöse System der Wirbeltiere (autonomes nervöses System) [The sympathetic and related nervous system of vertebrates (autonomic nervous system)]. Ergeb. Physiol., 27/II, 818–827.CrossRefGoogle Scholar
Langley, J. N. (1903b). The autonomic nervous system. Brain, 26, 1–26.CrossRefGoogle Scholar
Langley, J. N. (1921). The Autonomic Nervous System. Part I. Cambridge: W. Heffer.Google Scholar
Langley, J. N. and Anderson, H. K. (1895a). On the innervation of the pelvic and adjoining viscera. Part I. The lower portion of the intestine. J. Physiol., 18, 67–105.CrossRefGoogle Scholar
Langley, J. N. and Anderson, H. K. (1895b). The innervation of the pelvic and adjoining viscera. Part II. The bladder. J. Physiol., 19, 71–84.CrossRefGoogle Scholar
Langley, J. N. and Anderson, H. K. (1895c). The innervation of the pelvic and adjoining viscera. Part III. The external generative organs. J. Physiol., 19, 85–121.Google Scholar
Langley, J. N. and Anderson, H. K. (1895d). The innervation of the pelvic and adjoining viscera. Part IV. The internal generative organs. J. Physiol., 19, 122–130.Google Scholar
Langley, J. N. and Sherrington, C. S. (1891). On pilo-motor nerves. J. Physiol., 12, 278–291.CrossRefGoogle Scholar
Larsen, P. J., Enquist, L. W. and Card, J. P. (1998). Characterization of the multisynaptic neuronal control of the rat pineal gland using viral transneuronal tracing. Eur. J. Neurosci., 10, 128–145.CrossRefGoogle ScholarPubMed
Lavidis, N. A. and Bennett, M. R. (1992). Probabilistic secretion of quanta from visualized sympathetic nerve varicosities in mouse vas deferens. J. Physiol., 454, 9–26.CrossRefGoogle ScholarPubMed
Lawson, S. N. (1996). Neurochemistry of cutaneous nociceptors. In Neurobiology of Nociceptors, ed. Belmonte, C. and Cervero, F.. Oxford, New York, Tokyo: Oxford University Press, pp. 72–91.CrossRefGoogle Scholar
Lawson, S. N. (2005). The peripheral sensory nervous system: dorsal root ganglion neurons. In Peripheral Neuropathy, 4th edn., ed. Dyck, P. and Thomas, P. K.. Amsterdam: W. B. Saunders, Elsevier, pp. 163–202.Google Scholar
Douarin, N. M. and Kalcheim, C. (1999). The Neural Crest, 2nd edn. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
LeDoux, J. E. (1996). The Emotional Brain. New York: Simon & Shuster.Google Scholar
Lee, B. Y., Karmakar, M. G., Herz, B. L. and Sturgill, R. A. (1995). Autonomic dysreflexia revisited. J. Spinal Cord. Med., 18, 75–87.CrossRefGoogle ScholarPubMed
Lee, L. Y. and Pisarri, T. E. (2001). Afferent properties and reflex functions of bronchopulmonary C-fibers. Respir. Physiol., 125, 47–65.CrossRefGoogle ScholarPubMed
Leite-Panissi, C. R., Coimbra, N. C. and Menescal-de-Oliveira, L. (2003). The cholinergic stimulation of the central amygdala modifying the tonic immobility response and antinociception in guinea pigs depends on the ventrolateral periaqueductal gray. Brain Res. Bull., 60, 167–178.CrossRefGoogle ScholarPubMed
Levenson, R. W. (1993). Autonomic nervous system differences among emotions. Psychol. Sci., 3, 23–27.CrossRefGoogle Scholar
Levenson, R. W., Ekman, P. and Friesen, M. V. (1990). Voluntary facial action generates emotion-specific autonomic nervous system activity. Psychophysiology, 27, 363–384.CrossRefGoogle ScholarPubMed
Levenson, R. W., Carstensen, L. L., Friesen, W. V. and Ekman, P. (1991). Emotion, physiology, and expression in old age. Psychol. Aging, 6, 28–35.CrossRefGoogle ScholarPubMed
Levenson, R. W., Ekman, P., Heider, K. and Friesen, W. V. (1992). Emotion and autonomic nervous system activity in the Minangkabau of West Sumatra. J. Personal. Soc. Psychol., 62, 972–988.CrossRefGoogle ScholarPubMed
Levine, J. D., Fields, H. L. and Basbaum, A. L. (1993). Peptides and the primary afferent nociceptor. J. Neurosci., 13, 2273–2286.CrossRefGoogle ScholarPubMed
Lew, M. J., Rivers, R. J. and Duling, B. R. (1989). Arteriolar smooth muscle responses are modulated by an intramural diffusion barrier. Am. J. Physiol., 257, H10–H16.Google ScholarPubMed
Lewin, G. R. and McMahon, S. B. (1993). Muscle afferents innervating skin form somatotopically appropriate connections in the adult rat dorsal horn. Eur. J. Neurosci., 5, 1083–1092.CrossRefGoogle ScholarPubMed
Lewin, G. R., Ritter, A. M. and Mendell, L. M. (1993). Nerve growth factor-induced hyperalgesia in the neonatal and adult rat. J. Neurosci., 13, 2136–2148.CrossRefGoogle ScholarPubMed
Lewin, G. R., Rueff, A. and Mendell, L. M. (1994). Peripheral and central mechanisms of nerve growth factor-induced hyperalgesia. Eur. J. Neurosci., 6, 1903–1912.CrossRefGoogle Scholar
Lewis, D. I. and Coote, J. H. (1995). Chemical mediators of spinal inhibition of rat sympathetic neurones on stimulation in the nucleus tractus solitarii. J. Physiol., 486, 483–494.CrossRefGoogle ScholarPubMed
Lewis, D. I. and Coote, J. H. (1996). Baroreceptor-induced inhibition of sympathetic neurons by γ-aminobutyric acid acting at a spinal site. Am. J. Physiol., 270, H1885–H1892.Google Scholar
Lewis, M. W., Hermann, G. E., Rogers, R. C. and Travagli, R. A. (2002). In vitro and in vivo analysis of the effects of corticotropin releasing factor on rat dorsal vagal complex. J. Physiol., 543, 135–146.CrossRefGoogle ScholarPubMed
Li, Y. W. and Dampney, R. A. (1994). Expression of Fos-like protein in brain following sustained hypertension and hypotension in conscious rabbits. Neuroscience, 61, 613–634.CrossRefGoogle ScholarPubMed
Li, Y. W., Gieroba, Z. J., McAllen, R. M. and Blessing, W. W. (1991). Neurons in rabbit caudal ventrolateral medulla inhibit bulbospinal barosensitive neurons in rostral medulla. Am. J. Physiol., 261, R44–R51.Google ScholarPubMed
Lichtman, J. W. (1977). The reorganization of synaptic connexions in the rat submandibular ganglion during post-natal development. J. Physiol., 273, 155–177.CrossRefGoogle ScholarPubMed
Lichtman, J. W., Purves, D. and Yip, J. W. (1979). On the purpose of selective innervation of guinea-pig superior cervical ganglion cells. J. Physiol., 292, 69–84.CrossRefGoogle ScholarPubMed
Lichtman, J. W., Purves, D. and Yip, J. W. (1980). Innervation of sympathetic neurons in the guinea-pig thoracic chain. J. Physiol., 298, 285–299.CrossRefGoogle ScholarPubMed
Liddell, E. G. T. (1960). The Discovery of Reflexes. Oxford: Clarendon Press.Google Scholar
Lindgren, I. and Olivecrona, H. (1947). Surgical treatment of angina pectoris. J. Neurosurg., 4, 19–39.CrossRefGoogle ScholarPubMed
Lindh, B., Lundberg, J. M. and Hökfelt, T. (1989). neuropeptide Y-, galanin-, vasoactive intestinal peptide/PHI-, calcitonin gene-related peptide- and substance P-immunoreactive neuronal subpopulations in the cat autonomic and sensory ganglia and their projections. Cell Tissue Res., 256, 259–273.CrossRefGoogle ScholarPubMed
Lindh, B., Risling, M., Remahl, S., Terenius, L. and Hökfelt, T. (1993). Peptide-immunoreactive neurons and nerve in lumbosacral sympathetic ganglia: selective elimination of a pathway-specific expression of immunoreactivities following sciatic nerve resection in kittens. Neuroscience, 55, 545–562.CrossRefGoogle ScholarPubMed
Lipski, J., Kanjhan, R., Kruszewska, B. and Rong, W. (1996). Properties of presympathetic neurones in the rostral ventrolateral medulla in the rat: an intracellular study “in vivo'. J. Physiol., 490, 729–744.CrossRefGoogle ScholarPubMed
Lipski, J., Kawai, Y., Qi, J., Comer, A. and Win, J. (1998). Whole cell patch-clamp study of putative vasomotor neurons isolated from the rostral ventrolateral medulla. Am. J. Physiol., 274, R1099–R1110.Google ScholarPubMed
Lipski, J., Lin, J., Teo, M. Y. and Wyk, M. (2002). The network vs. pacemaker theory of the activity of RVL presympathetic neurons – a comparison with another putative pacemaker system. Auton. Neurosci., 98, 85–89.CrossRefGoogle Scholar
Llewellyn-Smith, I. J. and Weaver, L. C. (2001). Changes in synaptic inputs to sympathetic preganglionic neurons after spinal cord injury. J. Comp. Neurol., 435, 226–240.CrossRefGoogle ScholarPubMed
Llewellyn-Smith, I. J., Phend, K. D., Minson, J. B., Pilowsky, P. M. and Chalmers, J. P. (1992). Glutamate-immunoreactive synapses on retrogradely-labelled sympathetic preganglionic neurons in rat thoracic spinal cord. Brain Res., 581, 67–80.CrossRefGoogle ScholarPubMed
Llewellyn-Smith, I. J., Minson, J. B., Pilowsky, P. M., Arnolda, L. F. and Chalmers, J. P. (1995). The one hundred percent hypothesis: glutamate or γ-aminobutyric acid in synapses on sympathetic preganglionic neurons. Clin. Exp. Hypertens., 17, 323–333.CrossRefGoogle ScholarPubMed
Llewellyn-Smith, I. J., Cassam, A. K., Krenz, N. R., Krassioukov, A. V. and Weaver, L. C. (1997). Glutamate- and γ-aminobutyric acid-immunoreactive synapses on sympathetic preganglionic neurons caudal to a spinal cord transection in rats. Neuroscience, 80, 1225–1235.CrossRefGoogle Scholar
Llewellyn-Smith, I. J., Arnolda, L. F., Pilowsky, P. M., Chalmers, J. P. and Minson, J. B. (1998). γ-aminobutyric acid- and glutamate-immunoreactive synapses on sympathetic preganglionic neurons projecting to the superior cervical ganglion. J. Auton. Nerv. Syst., 71, 96–110.CrossRefGoogle ScholarPubMed
Lloyd, D. P. C. (1960). Spinal mechanisms involved in somatic activities. In Neurophysiology, Vol. II, ed. Field, J., Magoun, H. W. and Hall, V. E.. Washington: American Physiological Society, pp. 929–949.Google Scholar
Loewy, A. D. (1990a). Central autonomic pathways. In Central Regulation of Autonomic Functions, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 88–103.Google Scholar
Loewy, A. D. (1990b). Autonomic control of the eye. In Central Regulation of Autonomic Functions, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 268–285.Google Scholar
Loewy, A. D. (1998). Viruses as transneuronal tracers for defining neural circuits. Neurosci. Biobehav. Rev., 22, 679–684.CrossRefGoogle ScholarPubMed
Loewy, A. D. and Burton, H. (1978). Nuclei of the solitary tract: efferent projections to the lower brain stem and spinal cord of the cat. J. Comp. Neurol., 181, 421–449.CrossRefGoogle ScholarPubMed
Loewy, A. D. and Haxhiu, M. A. (1993). central nervous system cell groups projecting to pancreatic parasympathetic preganglionic neurons. Brain Res., 620, 323–330.CrossRefGoogle ScholarPubMed
Loewy, A. D. and Spyer, K. M. (1990a). Vagal preganglionic neurons. In Central Regulation of Autonomic Functions, ed. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 68–87.Google Scholar
Loewy, A. D. and Spyer, K. M. (eds.) (1990b). Central Regulation of Autonomic Functions. New York, Oxford: Oxford University Press.Google Scholar
Loewy, A. D., Franklin, M. F. and Haxhiu, M. A. (1994). central nervous system monoamine cell groups projecting to pancreatic vagal motor neurons: a transneuronal labeling study using pseudorabies virus. Brain Res., 638, 248–260.CrossRefGoogle ScholarPubMed
Löfving, B. (1961). Cardiovascular adjustments induced from the rostral cingulate gyrus with special reference to sympatho-inhibitory mechanisms. Acta Physiol. Scand., 53 (Suppl184), 1–82.Google ScholarPubMed
Logan, S. D., Pickering, A. E., Gibson, I. C., Nolan, M. F. and Spanswick, D. (1996). Electrotonic coupling between rat sympathetic preganglionic neurones in vitro. J. Physiol., 495, 491–502.CrossRefGoogle ScholarPubMed
Lombardi, F., Della Bella, P., Casati, R. and Malliani, A. (1981). Effects of intracoronary administration of bradykinin on the impulse activity of afferent sympathetic unmyelinated fibers with left ventricular endings in the cat. Circ. Res., 48, 69–75.CrossRefGoogle ScholarPubMed
Longhurst, J. C. (1995). Chemosensitive abdominal visceral afferents. In Visceral Pain. Progress in Pain Research and Management, Vol. 5, ed. Gebhart, G. F.. Seattle: IASP Press, pp. 99–132.Google Scholar
Lovén, C. (1866). Über die Erweiterung von Arterien in Folge einer Nervenerregung [On the vasodilation of arteries as a consequence of nerve stimulation]. Ber. Verh. königl. -sächs. Ges. Wiss.: Math. -phys. Classe, 18, 85–110.Google Scholar
Lovick, T. A. (1987). Differential control of cardiac and vasomotor activity by neurons in nucleus paragigantocellularis lateralis in the cat. J. Physiol., 389, 23–35.CrossRefGoogle ScholarPubMed
Low, P. (ed.) (1993). Clinical Autonomic Disorders. Boston, Toronto, London: Brown and Company.Google Scholar
Luckensmeyer, G. B. and Keast, J. R. (1998a). Characterisation of the adventitial rectal ganglia in the male rat by their immunohistochemical features and projections. J. Comp. Neurol., 396, 429–441.3.0.CO;2-3>CrossRefGoogle Scholar
Luckensmeyer, G. B. and Keast, J. R. (1998b). Projections of pelvic autonomic neurons within the lower bowel of the male rat: an anterograde labelling study. Neuroscience, 84, 263–280.CrossRefGoogle Scholar
Luff, S. E. and McLachlan, E. M. (1989). Frequency of neuromuscular junctions on arteries of different dimensions in the rabbit, guinea pig and rat. Blood Vessels, 26, 95–106.Google ScholarPubMed
Luff, S. E., McLachlan, E. M. and Hirst, G. D. (1987). An ultrastructural analysis of the sympathetic neuromuscular junctions on arterioles of the submucosa of the guinea pig ileum. J. Comp. Neurol., 257, 578–594.CrossRefGoogle ScholarPubMed
Luff, S. E., Hengstberger, S. G., McLachlan, E. M. and Anderson, W. P. (1991). Two types of sympathetic axon innervating the juxtaglomerular arterioles of the rabbit and rat kidney differ structurally from those supplying other arteries. J. Neurocytol., 20, 781–795.CrossRefGoogle ScholarPubMed
Luff, S. E., Hengstberger, S. G., McLachlan, E. M. and Anderson, W. P. (1992). Distribution of sympathetic neuroeffector junctions in the juxtaglomerular region of the rabbit kidney. J Auton. Nerv. Syst., 40, 239–253.CrossRefGoogle ScholarPubMed
Luff, S. E., Young, S. B. and McLachlan, E. M. (1995). Proportions and structure of contacting and non-contacting varicosities in the perivascular plexus of the rat tail artery. J. Comp. Neurol., 361, 699–709.CrossRefGoogle ScholarPubMed
Luff, S. E., Young, S. B. and McLachlan, E. M. (2000). Ultrastructure of substance P-immunoreactive terminals and their relation to vascular smooth muscle cells of rat small mesenteric arteries. J. Comp. Neurol., 416, 277–290.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Lundberg, A. (1971). Function of the ventral spinocerebellar tract. A new hypothesis. Exp. Brain Res., 12, 317–330.Google ScholarPubMed
Lundberg, A. (1975). Control of spinal mechansims from the brain. In The Basic Neurosciences, ed. Brady, R. C.. New York: Raven, pp. 253–265.Google Scholar
Lundberg, A. (1979). Multisensory control of spinal reflex pathways. Prog. Brain Res., 50, 11–28.CrossRefGoogle ScholarPubMed
Lundberg, J. M. (1981). Evidence for coexistence of vasoactive intestinal polypeptide (vasoactive intestinal peptide) and acetylcholine in neurons of cat exocrine glands. Morphological, biochemical and functional studies. Acta Physiol. Scand., 496, 1–57.Google ScholarPubMed
Lundberg, J. M. (1996). Pharmacology of cotransmission in the autonomic nervous system: integrative aspects on amines, neuropeptides, adenosine triphosphate, amino acids and nitric oxide. Pharmacol. Rev., 48, 113–178.Google ScholarPubMed
Lundberg, J. M., Hemsen, A., Rudehill, A.et al. (1988). Neuropeptide Y- and alpha-adrenergic receptors in pig spleen: localization, binding characteristics, cyclic AMP effects and functional responses in control and denervated animals. Neuroscience, 24, 659–672.CrossRefGoogle ScholarPubMed
Lundgren, O. (1988). Nervous control of intestinal transport. In Bailliere's Clinical Gastroenterology, Vol. 2/1, Gastrointestinal Neurophysiology, ed. Grundy, D. and Read, N. W.. London: Balliere Tindall, pp. 85–106.Google Scholar
Lundgren, O. (1989). Enteric nervous control of mucosal functions of the small intestine in vivo. In Nerves and the Gastrointestinal Tract, ed. Singer, M. V. and Goebell, A.. Dordrecht, The Netherlands: Kluwer Academic Publishers, pp. 275–285.Google Scholar
Lundgren, O. (2000). Sympathetic input into the enteric nervous system. Gut, 47, Suppl. 4, iv33–iv35.CrossRefGoogle ScholarPubMed
Luo, M., Hess, M. C., Fink, G. D.et al. (2003). Differential alterations in sympathetic neurotransmission in mesenteric arteries and veins in DOCA-salt hypertensive rats. Auton. Neurosci., 104, 47–57.CrossRefGoogle ScholarPubMed
Lykken, D. T. (1998). A Tremor in the Blood. New York: Plenum Press.Google Scholar
Lynn, B. (1996a). Neurogenic inflammation caused by cutaneous polymodal receptors. Prog. Brain Res., 113, 361–368.CrossRefGoogle Scholar
Lynn, B. (1996b). Efferent function of nociceptors. In Neurobiology of Nociceptors, ed. Belmonte, C. and Cervero, F.. Oxford, New York, Tokyo: Oxford University Press, pp. 418–438.CrossRefGoogle Scholar
Lynn, B., Schütterle, S. and Pierau, F. K. (1996). The vasodilator component of neurogenic inflammation is caused by a special subclass of heat-sensitive nociceptors in the skin of the pig. J. Physiol., 494, 587–593.CrossRefGoogle ScholarPubMed
Lynn, P. A. and Blackshaw, L. A. (1999). In vitro recordings of afferent fibres with receptive fields in the serosa, muscle and mucosa of rat colon. J. Physiol., 518, 271–282.CrossRefGoogle ScholarPubMed
Lynn, P. A., Olsson, C., Zagorodnyuk, V., Costa, M. and Brookes, S. J. (2003). Rectal intraganglionic laminar endings are transduction sites of extrinsic mechanoreceptors in the guinea pig rectum. Gastroenterology, 125, 786–794.CrossRefGoogle ScholarPubMed
Macefield, V. G. and Wallin, B. G. (1995). Modulation of muscle sympathetic activity during spontaneous and artificial ventilation and apnoea in humans. J. Auton. Nerv. Syst., 53, 137–147.CrossRefGoogle ScholarPubMed
Macefield, V. G. and Wallin, B. G. (1996). The discharge behaviour of single sympathetic neurones supplying human sweat glands. J. Physiol., 61, 277–286.Google ScholarPubMed
Macefield, V. G. and Wallin, B. G. (1999a). Firing properties of single vasoconstrictor neurones in human subjects with high level of muscle sympathetic activity. J. Physiol., 516, 293–301.CrossRefGoogle Scholar
Macefield, V. G. and Wallin, B. G. (1999b). Respiratory and cardiac modulation of single sympathetic vasoconstrictor and sudomotor neurones to human skin. J. Physiol., 516, 303–314.CrossRefGoogle Scholar
Macefield, V. G., Wallin, B. G. and Vallbo, A. B. (1994). The discharge behaviour of single vasoconstrictor motoneurones in human muscle nerves. J. Physiol., 481, 799–809.CrossRefGoogle ScholarPubMed
Macefield, V. G., Elam, M. and Wallin, B. G. (2002). Firing properties of single postganglionic sympathetic neurons recorded in awake human subjects. Auton. Neurosci., 95, 146–159.CrossRefGoogle ScholarPubMed
Macefield, V. G., Sverrisdottir, Y. B. and Wallin, B. G. (2003). Resting discharge of human muscle spindles is not modulated by increases in sympathetic drive. J. Physiol., 551, 1005–1011.CrossRefGoogle Scholar
Madden, K. S. and Felten, D. L. (1995). Experimental basis for neural-immune interactions. Physiol. Rev., 75, 77–106.CrossRefGoogle ScholarPubMed
Madden, C. J. and Sved, A. F. (2003). Cardiovascular regulation after destruction of the C1 cell group of the rostral ventrolateral medulla in rats. Am. J. Physiol. Heart Circ. Physiol., 285, H2734–H2748.CrossRefGoogle ScholarPubMed
Madden, K. S., Sanders, V. M. and Felten, D. L. (1995). Catecholamine influences and sympathetic modulation of immune responsiveness. Rev. Pharmacol. Toxicol., 35, 417–448.CrossRefGoogle ScholarPubMed
Maggi, C. A. (ed.) (1993). The Autonomic Nervous System, Vol. 3, Nervous Control of the Urogenital System. Chur, Switzerland: Harwood Academic Publishers.Google Scholar
Maggi, C. A. and Meli, A. (1988). The sensory-efferent function of capsaicin-sensitive sensory neurons. Gen. Pharmacol., 19, 1–43.CrossRefGoogle ScholarPubMed
Maggi, C. A., Giachetti, A., Dey, R. D. and Said, S. I. (1995). Neuropeptides as regulators for airway function: vasoactive intestinal peptide and the tachykinins. Physiol. Rev., 75, 277–322.CrossRefGoogle ScholarPubMed
Maier, S. F. and Watkins, L. R. (1998). Cytokines for psychologists: implications of bidirectional immune-to-brain communication for understanding behavior, mood, and cognition. Psychol. Rev., 105, 83–107.CrossRefGoogle Scholar
Malliani, A. (1982). Cardiovascular sympathetic afferent fibers. Rev. Physiol. Biochem. Pharmacol., 94, 11–74.CrossRefGoogle Scholar
Malliani, A. and Lombardi, F. (1982). Consideration of the fundamental mechanisms eliciting cardiac pain. Am. Heart J., 103, 575–578.CrossRefGoogle ScholarPubMed
Malpas, S. C. (1998). The rhythmicity of sympathetic nerve activity. Prog. Neurobiol., 56, 65–96.CrossRefGoogle ScholarPubMed
Mancia, G., Baccelli, G. and Zanchetti, A. (1972). Hemodynamic responses to different emotional stimuli in the cat: patterns and mechanisms. Am. J. Physiol., 223, 925–933.Google ScholarPubMed
Mano, T. (1999). Muscular and cutaneous sympathetic activity. In Handbook of Clinical Neurology, Vol. 74, The Autonomic Nervous System, part I: Normal Functions, ed. Appenzeller, O.. Amsterdam: Elsevier, pp. 649–665.Google Scholar
Mantyh, P. W., Rogers, S. D., Honore, P.et al. (1997). Inhibition of hyperalgesia by ablation of lamina I spinal neurons expressing the substance P receptor. Science, 278, 275–279.CrossRefGoogle ScholarPubMed
Markakis, E. A. and Swanson, L. W. (1997). Spatiotemporal patterns of secretomotor neuron generation in the parvicellular neuroendocrine system. Brain Res. Brain Res. Rev., 24, 255–291.CrossRefGoogle ScholarPubMed
Marsh, D. R. and Weaver, L. C. (2004). Autonomic dysreflexia, induced by noxious or innocuous stimulation, does not depend on changes in dorsal horn substance p. J. Neurotrauma, 21, 817–828.CrossRefGoogle Scholar
Marsh, D. R., Wong, S. T., Meakin, S. O.et al. (2002). Neutralizing intraspinal nerve growth factor with a trkA-IgG fusion protein blocks the development of autonomic dysreflexia in a clip-compression model of spinal cord injury. J. Neurotrauma, 19, 1531–1541.CrossRefGoogle Scholar
Marshall, J. M. (1994). Peripheral chemoreceptors and cardiovascular regulation. Physiol. Rev., 74, 543–594.CrossRefGoogle ScholarPubMed
Marson, L. (1995). Central nervous system neurons identified after injection of pseudorabies virus into the rat clitoris. Neurosci. Lett., 190, 41–44.CrossRefGoogle ScholarPubMed
Marson, L. (1997). Identification of central nervous system neurons that innervate the bladder body, bladder base, or external urethral sphincter of female rats: a transneuronal tracing study using pseudorabies virus. J. Comp. Neurol., 389, 584–602.3.0.CO;2-X>CrossRefGoogle ScholarPubMed
Marson, L. and McKenna, K. E. (1996). central nervous system cell groups involved in the control of the ischiocavernosus and bulbospongiosus muscles: a transneuronal tracing study using pseudorabies virus. J. Comp. Neurol., 374, 161–179.3.0.CO;2-0>CrossRefGoogle ScholarPubMed
Marson, L., Platt, K. B. and McKenna, K. E. (1993). Central nervous system innervation of the penis as revealed by the transneuronal transport of pseudorabies virus. Neuroscience, 55, 263–280.CrossRefGoogle ScholarPubMed
Martinez-Pena y Valencuela, I., Rogers, R. C., Hermann, G. E. and Travagli, R. A. (2004). Norepinephrine effects on identified neurons of the rat dorsal motor nucleus of the vagus. Am. J. Physiol. Gastrointest. Liver Physiol., 286, G333–G339.Google Scholar
Martínez-Pinna, J., Davies, P. J. and McLachlan, E. M. (2000). Diversity of channels involved in Ca2 + activation of K+ channels during the prolonged AHP in guinea-pig sympathetic neurons. J. Neurophysiol., 84, 1346–1354.CrossRefGoogle Scholar
Mason, P. (2001). Contributions of the medullary raphe and ventromedial reticular region to pain modulation and other homeostatic functions. Annu. Rev. Neurosci., 24, 737–777.CrossRefGoogle ScholarPubMed
Mathias, C. J. and Bannister, R. (eds.) (2002). Autonomic Failure, 4th edn. Oxford: Oxford University Press.Google ScholarPubMed
Mathias, C. J. and Frankel, H. L. (2002). Autonomic disturbances and spinal cord lesions. In Autonomic Failure, 4th edn., ed. Mathias, C. J. and Bannister, R.. Oxford: Oxford University Press, pp. 494–513.Google ScholarPubMed
Matsuo, R. and Kang, Y. (1998). Two types of parasympathetic preganglionic neurones in the superior salivatory nucleus characterized electrophysiologically in slice preparations of neonatal rats. J. Physiol., 513, 157–170.CrossRefGoogle ScholarPubMed
Matsuo, R. and Yamamoto, T. (1989). Gustatory-salivary reflex: neural activity of sympathetic and parasympathetic fibers innervating the submandibular gland of the hamster. J. Auton. Nerv. Syst., 26, 187–197.CrossRefGoogle ScholarPubMed
Matsuo, R., Morimoto, T. and Kang, Y. (1998). Neural activity of the superior salivatory nucleus in rats. Eur. J. Morphol., 36 Suppl, 203–207.Google ScholarPubMed
Matthews, L. H. (1969). The Life of Mammals. London: Weidenfeld and Nicolson.Google Scholar
Matthews, M. R. and Cuello, A. C. (1984). The origin and possible significance of substance P immunoreactive networks in the prevertebral ganglia and related structures in the guinea-pig. Philos. Trans. R. Soc. London B Biol. Sci., 306, 247–276.CrossRefGoogle ScholarPubMed
Matthews, M. R., Connaughton, M. and Cuello, A. C. (1987). Ultrastructure and distribution of substance P-immunoreactive sensory collaterals in the guinea pig prevertebral sympathetic ganglia. J. Comp. Neurol., 258, 28–51.CrossRefGoogle ScholarPubMed
Mawe, G. M. (1995). Prevertebral, pancreatic and gallbladder ganglia: non-enteric ganglia that are involved in gastrointestinal function. In The Autonomic Nervous System, Vol. 6, Autonomic Ganglia, ed. McLachlan, E. M.. Luxembourg: Harwood Academic Publishers, pp. 397–444.Google Scholar
Mawe, G. M. (1998). Nerves and hormones interact to control gallbladder function. News Physiol. Sci., 13, 84–90.Google ScholarPubMed
Mayer, E. A. and Raybould, H. E. (eds.) (1993). Basic and Clinical Aspects of Chronic Abdominal Pain. Amsterdam: Elsevier Science Publishers B.V.Google Scholar
Mayer, E. A., Munakata, J., Mertz, H., Lembo, T. and Bernstein, C. N. (1995). Visceral hyperalgesia and irritable bowel syndrome. In Visceral Pain, ed. Gebhart, G. F.. Seattle: IASP Press, pp. 429–468.Google Scholar
McAllen, R. M. (1987). Central respiratory modulation of subretrofacial bulbospinal neurones in the cat. J. Physiol., 388, 533–545.CrossRefGoogle ScholarPubMed
McAllen, R. M. (1992). Actions of carotid chemoreceptors on subretrofacial bulbospinal neurons in the cat. J. Auton. Nerv. Syst., 40, 181–188.CrossRefGoogle ScholarPubMed
McAllen, R. M. and May, C. N. (1994a). Differential drives from rostral ventrolateral medullary neurons to three identified sympathetic outflows. Am. J. Physiol., 267, R935–R944.Google Scholar
McAllen, R. M. and May, C. N. (1994b). Effects of preoptic warming on subretrofacial and cutaneous vasoconstrictor neurons in anaesthetized cats. J. Physiol., 481, 719–730.CrossRefGoogle Scholar
McAllen, R. M. and Spyer, K. M. (1978a). Two types of vagal preganglionic motoneurones projecting to the heart and lungs. J. Physiol., 282, 353–364.CrossRefGoogle Scholar
McAllen, R. M. and Spyer, K. M. (1978b). The baroreceptor input to cardiac vagal motoneurones. J. Physiol., 282, 365–374.CrossRefGoogle Scholar
McAllen, R. M., Häbler, H. J., Michaelis, M., Peters, O. and Jänig, W. (1994). Monosynaptic excitation of preganglionic vasomotor neurons by subretrofacial neurons of the rostral ventrolateral medulla. Brain Res., 634, 227–234.CrossRefGoogle ScholarPubMed
McAllen, R. M., May, C. N. and Shafton, A. D. (1995). Functional anatomy of sympathetic premotor cell groups in the medulla. Clin. Exp. Hypertens., 17, 209–221.CrossRefGoogle ScholarPubMed
McAllen, R. M., May, C. N. and Campos, R. R. (1997). The supply of vasomotor drive to individual classes of sympathetic neuron. Clin. Exp. Hypertens., 19, 607–618.CrossRefGoogle ScholarPubMed
McCabe, P. M., Duan, Y. F., Winters, R. W.et al. (1994). Comparison of peripheral blood flow patterns associated with the defense reaction and the vigilance reaction in rabbits. Physiol. Behav., 56, 1101–1106.CrossRefGoogle ScholarPubMed
McCall, R. B., Gebber, G. L. and Barman, S. M. (1977). Spinal interneurons in the baroreceptor reflex arc. Am. J. Physiol., 232, H657–H665.Google ScholarPubMed
McCrea, D. A. (1994). Can sense be made of spinal interneuron circuits? In Movement Control, ed. Cordo, P. and Harnad, S.. Cambridge: Cambridge University Press, pp. 31–41.CrossRefGoogle Scholar
McCrea, D. A. (2001). Spinal circuitry of sensorimotor control of locomotion. J. Physiol., 533, 41–50.CrossRefGoogle ScholarPubMed
McDonald, D. M. (1990). The ultrastructure and permeability of tracheobronchial blood vessels in health and disease. Eur. Respir. J. Suppl., 12, 572s–585s.Google ScholarPubMed
McDonald, D. M. (1997). Neurogenic inflammation in the airways. In The Autonomic Nervous System, Vol. 7, Autonomic Control of the Respiratory System, ed. Barnes, P. J.. Amsterdam: Harwood Academic Publishers GmbH, pp. 249–289.Google Scholar
McDonald, D. M., Mitchell, R. A., Gabella, G. and Haskell, A. (1988). Neurogenic inflammation in the rat trachea. II. Identity and distribution of nerves mediating the increase in vascular permeability. J. Neurocytol., 17, 605–628.CrossRefGoogle ScholarPubMed
McEwen, B. S. (1998). Protective and damaging effects of stress mediators. New Engl. J. Med., 338, 171–179.CrossRefGoogle ScholarPubMed
McEwen, B. S. (2000). Protective and damaging effects of stress mediators: central role of the brain. Prog. Brain Res., 122, 25–34.CrossRefGoogle Scholar
McEwen, B. S. (ed.) (2001a). Handbook of Physiology. Section 7: The Endocrine System, Vol. IV, Coping with the Environment: Neural and Neuroendocrine Mechanisms. Oxford, New York: Oxford University Press.Google Scholar
McEwen, B. S. (2001b). Neurobiology of interpreting and responding to stressful events: paradigmatic role of the hippocampus. In Handbook of Physiology. Section 7: The Endocrine System, Vol. IV, Coping with the Environment: Neural and Neuroendocrine Mechanisms, ed. McEwen, B. S., pp. 155–178. Oxford University Press, Oxford New York.Google Scholar
McEwen, B. S. and Wingfield, J. C. (2003). The concept of allostasis in biology and biomedicine. Horm. Behav., 43, 2–15.CrossRefGoogle ScholarPubMed
McGaugh, J. L. (2000). Memory – a century of consolidation. Science, 287, 248–251.CrossRefGoogle ScholarPubMed
McGaugh, J. L. and Roozendaal, B. (2002). Role of adrenal stress hormones in forming lasting memories in the brain. Curr. Opin. Neurobiol., 12, 205–210.CrossRefGoogle Scholar
McKenna, K. E. (1999). Central nervous system pathways involved in the control of penile erection. Annu. Rev. Sex Res., 10, 157–183.Google ScholarPubMed
McKenna, K. E. (2000). The neural control of female sexual function. NeuroRehabilitation, 15, 133–143.Google ScholarPubMed
McKenna, K. E. (2001). Neural circuitry involved in sexual function. J. Spinal Cord Med., 24, 148–154.CrossRefGoogle ScholarPubMed
McKenna, K. E. (2002). The neurophysiology of female sexual function. World J. Urol., 20, 93–100.CrossRefGoogle ScholarPubMed
McKenna, K. E. and Marson, L. (1997). Spinal and brain stem control of sexual function. In The Autonomic Nervous System, Vol. 11, Central Nervous Control of Autonomic Function, ed. Jordan, D.. Amsterdam: Harwood Academic Publishers, pp. 151–187.Google Scholar
McKenna, K. E., Chung, S. K. and McVary, K. T. (1991). A model for the study of sexual function in anesthetized male and female rats. Am. J. Physiol., 261, R1276–R1285.Google Scholar
McKinley, M. J., Clarke, I. J. and Oldfield, B. J. (2004). Circumventricular organs. In The Human Nervous System, ed. Paxinos, G. and Mai, J. K.. Amsterdam: Elsevier Academic Press, pp. 562–591.Google Scholar
McLachlan, E. M. (1975). An analysis of the release of acetylcholine from preganglionic nerve terminals. J. Physiol., 245, 447–466.CrossRefGoogle ScholarPubMed
McLachlan, E. M. (1985). The components of the hypogastric nerve in male and female guinea pigs. J. Auton. Nerv. Syst., 13, 327–342.CrossRefGoogle ScholarPubMed
McLachlan, E. M. (ed.) (1995). The Autonomic Nervous System, Vol. 6, Autonomic Ganglia. Luxembourg: Harwood Academic Publishers.Google Scholar
McLachlan, E. M. and Hirst, G. D. (1980). Some properties of preganglionic neurons in upper thoracic spinal cord of the cat. J. Neurophysiol., 43, 1251–1265.CrossRefGoogle ScholarPubMed
McLachlan, E. M. and Jänig, W. (1983). The cell bodies of origin of sympathetic and sensory axons in some skin and muscle nerves of the cat hindlimb. J. Comp. Neurol., 214, 115–130.CrossRefGoogle ScholarPubMed
McLachlan, E. M. and Meckler, R. L. (1989). Characteristics of synaptic input to three classes of sympathetic neurone in the coeliac ganglion of the guinea-pig. J. Physiol., 415, 109–129.CrossRefGoogle ScholarPubMed
McLachlan, E. M., Davies, P. J., Häbler, H. J. and Jamieson, J. (1997). Ongoing and reflex synaptic events in rat superior cervical ganglion cells. J. Physiol., 501, 165–181.CrossRefGoogle Scholar
McLachlan, E. M., Häbler, H. J., Jamieson, J. and Davies, P. J. (1998). Analysis of the periodicity of synaptic events in neurones in the superior cervical ganglion of anaesthetized rats. J. Physiol., 511, 461–478.CrossRefGoogle ScholarPubMed
McMahon, S. B. (1996). nerve growth factor as a mediator of inflammatory pain. Philos. Trans. R. Soc. London B Biol. Sci., 351, 431–440.CrossRefGoogle ScholarPubMed
McMahon, S. B. and Morrison, J. F. (1982a). Spinal neurones with long projections activated from the abdominal viscera of the cat. J. Physiol., 322, 1–20.CrossRefGoogle Scholar
McMahon, S. B. and Morrison, J. F. (1982b). Two groups of spinal interneurones that respond to stimulation of the abdominal viscera of the cat. J. Physiol., 322, 21–34.CrossRefGoogle Scholar
Meckler, R. L. and Weaver, L. C. (1988). Characteristics of ongoing and reflex discharge of single splenic and renal sympathetic postganglionic fibres in the cat. J. Physiol., 396, 139–153.CrossRefGoogle Scholar
Mei, N. (1983). Sensory structures in the viscera. Prog. Sensory Physiol., 4, 1–42.CrossRefGoogle Scholar
Mei, N. (1985). Intestinal chemosensitivity. Physiol. Rev., 65, 211–237.CrossRefGoogle ScholarPubMed
Meller, S. T. and Gebhart, G. F. (1992). A critical review of the afferent pathways and the potential chemical mediators involved in cardiac pain. Neuroscience, 48, 501–524.CrossRefGoogle ScholarPubMed
Melnitchenko, L. V. and Skok, V. I. (1970). Natural electrical activity in mammalian parasympathetic ganglion neurones. Brain Res., 23, 277–279.CrossRefGoogle ScholarPubMed
Menendez, L., Bester, H., Besson, J. M. and Bernard, J. F. (1996). Parabrachial area: electrophysiological evidence for an involvement in cold nociception. J. Neurophysiol., 75, 2099–2116.CrossRefGoogle ScholarPubMed
Mense, S. and Craig, A. D. Jr. (1988). Spinal and supraspinal terminations of primary afferent fibers from the gastrocnemius-soleus muscle in the cat. Neuroscience, 26, 1023–1035.CrossRefGoogle ScholarPubMed
Messenger, J. P., Anderson, R. L. and Gibbins, I. L. (1999). Neurokinin-1 receptor localisation in guinea pig autonomic ganglia. J. Comp. Neurol., 412, 693–704.3.0.CO;2-T>CrossRefGoogle ScholarPubMed
Meyer, R. A., Davis, K. D., Cohen, R. H., Treede, R. D. and Campbell, J. N. (1991). Mechanically insensitive afferents (MIAs) in cutaneous nerves of monkey. Brain Res., 561, 252–261.CrossRefGoogle ScholarPubMed
Meyer, R. A., Ringkamp, M., Campbell, J. N. and Raja, S. N. (2005). Peripheral mechanisms of cutaneous nociception. In Wall and Mezack's Textbook of Pain, 5th edn., ed. McMahon, S. B. and Koltzenburg, M.. Edinburgh: Elsevier Churchill Livingstone, pp. 3–34.Google Scholar
Meyers, G. E. (1986). William James, His Life and Thought. New Haven: Yale University Press.Google Scholar
Miao, F. J. P., Green, P. G., Coderre, T. J., Jänig, W. and Levine, J. D. (1996a). Sympathetic-dependence in bradykinin-induced synovial plasma extravasation is dose-related. Neurosci. Lett., 205, 165–168.CrossRefGoogle Scholar
Miao, F. J. P., Jänig, W. and Levine, J. D. (1996b). Role of sympathetic postganglionic neurons in synovial plasma extravasation induced by bradykinin. J. Neurophysiol., 75, 715–724.CrossRefGoogle Scholar
Miao, F. J. P., Jänig, W., Green, P. G. and Levine, J. D. (1997a). Inhibition of bradykinin-induced synovial plasma extravasation produced by noxious cutaneous and visceral stimuli and its modulation by activity in the vagal nerve. J. Neurophysiol., 78, 1285–1292.CrossRefGoogle Scholar
Miao, F. J. P., Jänig, W. and Levine, J. D. (1997b). Vagal branches involved in inhibition of bradykinin-induced synovial plasma extravasation by intrathecal nicotine and noxious stimulation in the rat. J. Physiol., 498, 473–481.CrossRefGoogle Scholar
Miao, F. J. P., Jänig, W. and Levine, J. D. (2000). Nociceptive neuroendocrine negative feedback control of neurogenic inflammation activated by capsaicin in the rat paw: role of the adrenal medulla. J. Physiol., 527, 601–610.CrossRefGoogle ScholarPubMed
Miao, F. J. P., Jänig, W., Jasmin, L. and Levine, J. D. (2001). Spino-bulbospinal pathway mediating vagal modulation of nociceptive-neuroendocrine control of inflammation in the rat. J. Physiol., 532, 811–822.CrossRefGoogle Scholar
Miao, F. J. P., Green, P. G. and Levine, J. D. (2003b). Mechano-sensitive duodenal afferents contribute to vagal modulation of inflammation in the rat. J. Physiol., 554, 227–235.CrossRefGoogle Scholar
Miao, F. J. P., Jänig, W., Jasmin, L. and Levine, J. D. (2003a). Blockade of nociceptive inhibition of plasma extravasation by opioid stimulation of the periaqueductal gray and its interaction with vagus-induced inhibition in the rat. Neuroscience, 119, 875–885.CrossRefGoogle Scholar
Michaelis, M., Göder, R., Häbler, H. J. and Jänig, W. (1994). Properties of afferent nerve fibres supplying the saphenous vein in the cat. J. Physiol., 474, 233–243.CrossRefGoogle ScholarPubMed
Michaelis, M., Häbler, H. J. and Jänig, W. (1996). Silent afferents: a separate class of primary afferents?Clin. Exp. Pharmacol. Physiol., 23, 99–105.CrossRefGoogle ScholarPubMed
Michl, T., Jocic, M., Heinemann, A., Schuligoi, R. and Holzer, P. (2001). Vagal afferent signaling of a gastric mucosal acid insult to medullary, pontine, thalamic, hypothalamic and limbic, but not cortical, nuclei of the rat brain. Pain, 92, 19–27.CrossRefGoogle Scholar
Mifflin, S. W., Spyer, K. M. and Withington-Wray, D. J. (1988). Baroreceptor inputs to the nucleus tractus solitarius in the cat: postsynaptic actions and the influence of respiration. J. Physiol., 399, 349–367.CrossRefGoogle Scholar
Milner, T. A., Morrison, S. F., Abate, C. and Reis, D. J. (1988). Phenylethanolamine N-methyltransferase-containing terminals synapse directly on sympathetic preganglionic neurons in the rat. Brain Res., 448, 205–222.CrossRefGoogle ScholarPubMed
Minson, J. B., Llewellyn-Smith, I. J., Chalmers, J. P., Pilowsky, P. M. and Arnolda, L. F. (1997). c-fos identifies γ-aminobutyric acid-synthesizing barosensitive neurons in caudal ventrolateral medulla. NeuroReport, 8, 3015–3021.CrossRefGoogle Scholar
Mitchell, J. H. (1985). Cardiovascular control during exercise: central and reflex neural mechanisms. Am. J. Cardiol., 55, 34D–41D.CrossRefGoogle ScholarPubMed
Mitchell, R. A., Herbert, D. A., Baker, D. G. and Basbaum, C. B. (1987). In vivo activity of tracheal parasympathetic ganglion cells innervating tracheal smooth muscle. Brain Res., 437, 157–160.CrossRefGoogle ScholarPubMed
Mitchell, S. W. (1872). Injuries of Nerves and their Consequences. Philadelphia: JB Lippincott.Google Scholar
Miura, M., Takayama, K. and Okada, J. (1994). Distribution of glutamate- and γ-aminobutyric acid-immunoreactive neurons projecting to the cardioacceleratory center of the intermediolateral nucleus of the thoracic cord of SHR and WKY rats: a double-labeling study. Brain Res., 638, 139–150.CrossRefGoogle Scholar
Miyawaki, T., Goodchild, A. K. and Pilowsky, P. M. (2003). Maintenance of sympathetic tone by a nickel chloride sensitive mechanism in the rostral ventrolateral medulla of the adult rat. Neuroscience, 116, 455–464.CrossRefGoogle ScholarPubMed
Molander, C. and Grant, G. (1995). Spinal cord cytoarchitecture. In The Rat Nervous System, ed. Paxinos, G.. San Diego: Academic Press, pp. 39–45.Google Scholar
Monnier, A., Alheid, G. F. and McCrimmon, D. R. (2003). Defining ventral medullary respiratory compartments with a glutamate receptor agonist in the rat. J. Physiol., 548, 859–874.CrossRefGoogle ScholarPubMed
Moore, R. Y. (1996). Neural control of the pineal gland. Behav. Brain Res., 73, 125–130.CrossRefGoogle ScholarPubMed
Moore, R. Y. (2003). Circadian timing. In Fundamental Neuroscience, 2nd edn., ed. Squire, L. R., Bloom, F. E., McConnell, S. K.et al. San Diego: Academic Press, pp. 1067–1084.Google Scholar
Moran, T. H. and Schwartz, G. J. (1994). Neurobiology of cholecystokinin. Crit. Rev. Neurobiol., 9, 1–28.Google ScholarPubMed
Morgan, C. W., Groat, W. C., Felkins, L. A. and Zhang, S. J. (1993). Intracellular injection of neurobiotin or horseradish peroxidase reveals separate types of preganglionic neurons in the sacral parasympathetic nucleus of the cat. J. Comp. Neurol., 331, 161–182.CrossRefGoogle ScholarPubMed
Moriarty, M., Gibbins, I. L., Potter, E. K. and McCloskey, D. I. (1992). Comparison of the inhibitory roles of neuropeptide Y and galanin on cardiac vagal action in the dog. Neurosci. Lett., 139, 275–279.CrossRefGoogle ScholarPubMed
Morris, J. L. (1995). Distribution and peptide content of sympathetic axons innervating different regions of the cutaneous venous bed in the pinna of the guinea pig ear. J. Vasc. Res., 32, 378–386.CrossRefGoogle ScholarPubMed
Morris, J. L. (1999). Cotransmission from sympathetic vasoconstrictor neurons to small cutaneous arteries in vivo. Am. J. Physiol., 277, H58–H64.Google ScholarPubMed
Morris, J. and Dolan, R. (2004). Functional neuroanatomy of human emotion. In Human Brain Function, 2nd edn., ed. Frackowiak, R. S. J., Friston, K. J., Frith, C. D.et al. Amsterdam: Elsevier Academic Press, pp. 365–396.Google Scholar
Morris, J. L. and Gibbins, I. L. (1992). Co-transmission and neuromodulation. In The Autonomic Nervous System, Vol. 1, Autonomic Neuroeffector Mechanisms, ed. Burnstock, G. and Hoyle, C. H. V.. Chur, Switzerland: Harwood, pp. 33–119.Google Scholar
Morris, J. L. and Gibbins, I. L. (eds.) (1997). The Autonomic Nervous System, Vol. 12, Autonomic Innervation of the Skin, Chur, Switzerland: Harwood Academic Publishers.Google Scholar
Morris, J. L., Gibbins, I. L. and Clevers, J. (1981). Resistance of adrenergic neurotransmission in the toad heart to adrenoceptor blockade. Naunyn-Schmiedeberg's Arch. Pharmacol., 317, 331–338.CrossRefGoogle ScholarPubMed
Morris, M. J., Russell, A. E., Kapoor, V.et al. (1986). Increases in plasma neuropeptide Y concentrations during sympathetic activation in man. J. Auton. Nerv. Syst., 17, 143–149.CrossRefGoogle ScholarPubMed
Morris, J. L., Gibbins, I. L. and Jobling, P. (2005). Post-stimulus potentiation of transmission in pelvic ganglia enhances sympathetic dilatation of guinea-pig uterine artery in vitro. J. Physiol., 566, 189–203.CrossRefGoogle ScholarPubMed
Morrison, J. F. B. (1997). Central nervous control of the bladder. In The Autonomic Nervous System, Vol. 11, Central Nervous Control of Autonomic Function, ed. Jordan, D.. Amsterdam: Harwood Academic Publishers, pp. 129–149.Google Scholar
Morrison, S. F. (1999). rostral ventrolateral medulla and raphe differentially regulate sympathetic outflows to splanchnic and brown adipose tissue. Am. J. Physiol., 276, R962–R973.Google ScholarPubMed
Morrison, S. F. (2001a). Differential control of sympathetic outflow. Am. J. Physiol. Regul. Integr. Comp. Physiol., 281, R683–R698.CrossRefGoogle Scholar
Morrison, S. F. (2001b). Differential regulation of brown adipose and splanchnic sympathetic outflows in rat: roles of raphe and rostral ventrolateral medulla neurons. Clin. Exp. Pharmacol. Physiol., 28, 138–143.CrossRefGoogle Scholar
Morrison, S. F. and Cao, W. H. (2000). Different adrenal sympathetic preganglionic neurons regulate epinephrine and norepinephrine secretion. Am. J. Physiol. Regul. Integr. Comp. Physiol., 279, R1763–R1775.CrossRefGoogle ScholarPubMed
Morrison, S. F. and Gebber, G. L. (1985). Axonal branching patterns and funicular trajectories of raphespinal sympathoinhibitory neurons. J. Neurophysiol., 53, 759–772.CrossRefGoogle ScholarPubMed
Morrison, S. F., Callaway, J., Milner, T. A. and Reis, D. J. (1991). Rostral ventrolateral medulla: a source of the glutamatergic innervation of the sympathetic intermediolateral nucleus. Brain Res., 562, 126–135.CrossRefGoogle ScholarPubMed
Morrison, S. F., Sved, A. F. and Passerin, A. M. (1999). γ-aminobutyric acid-mediated inhibition of raphe pallidus neurons regulates sympathetic outflow to brown adipose tissue. Am. J. Physiol., 276, R290–R297.Google Scholar
Mosqueda-Garcia, R., Furlan, R., Tank, J. and Fernandez-Violante, R. (2000). The elusive pathophysiology of neurally mediated syncope. Circulation, 102, 2898–2906.CrossRefGoogle ScholarPubMed
Müller, L. R. (1902). Klinische und experimentelle Studien ueber die Innervation der Blase, des Mastdarmes und des Genitalapparates [Clinical and experimental studies on the innervation of the urinary bladder, hindgut and reproductive organs]. Dtsch. Z. Nervenheilkd., 21, 86–155.CrossRefGoogle Scholar
Müller, L. R. (1906). Ueber die Exstirpation der unteren Haelfte des Rueckenmarkes und deren Folgeerscheinungen [On the consequences of the removal of the lower half of the spinal cord]. Dtsch. Z. Nervenheilkd., 30, 411–423.CrossRefGoogle Scholar
Mulryan, K., Gitterman, D. P., Lewis, C. J.et al. (2000). Reduced vas deferens contraction and male infertility in mice lacking P2X1 receptors. Nature, 403, 86–89.CrossRefGoogle ScholarPubMed
Mulvany, M. J., Nilsson, H. and Flatman, J. A. (1982). Role of membrane potential in the response of rat small mesenteric arteries to exogenous noradrenaline stimulation. J. Physiol., 332, 363–373.CrossRefGoogle ScholarPubMed
Murphy, S. M., Matthew, S. E., Rodgers, H. F., Lituri, D. T. and Gibbins, I. L. (1998). Synaptic organisation of neuropeptide-containing preganglionic boutons in lumbar sympathetic ganglia of guinea pigs. J. Comp. Neurol., 398, 551–567.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Murray, J. G. and Thompson, J. W. (1957). The occurrence and function of collateral sprouting in the sympathetic nervous system of the cat. J. Physiol., 135, 133–162.CrossRefGoogle ScholarPubMed
Nadelhaft, I. and Vera, P. L. (1995). Central nervous system neurons infected by pseudorabies virus injected into the rat urinary bladder following unilateral transection of the pelvic nerve. J. Comp. Neurol., 359, 443–456.CrossRefGoogle ScholarPubMed
Nadelhaft, I., Vera, P. L., Card, J. P. and Miselis, R. R. (1992). Central nervous system neurons labelled following the injection of pseudorabies virus into the rat urinary bladder. Neurosci. Lett., 143, 271–274.CrossRefGoogle ScholarPubMed
Nagashima, K., Nakai, S., Tanaka, M. and Kanosue, K. (2000). Neuronal circuitries involved in thermoregulation. Auton. Neurosci., 85, 18–25.CrossRefGoogle ScholarPubMed
Natarajan, M. and Morrison, S. F. (2000). Sympathoexcitatory caudal ventrolateral medulla neurons mediate responses to caudal pressor area stimulation. Am. J. Physiol. Regul. Integr. Comp. Physiol., 279, R364–R374.CrossRefGoogle ScholarPubMed
Neff, R. A., Mihalevich, M. and Mendelowitz, D. (1998). Stimulation of nucleus tractus solitarii activates N-methyl-D-aspartate (acid) and non-N-methyl-D-aspartate (acid) receptors in rat cardiac vagal neurons in the nucleus ambiguus. Brain Res., 792, 277–282.CrossRefGoogle Scholar
Ness, T. J. and Gebhart, G. F. (1990). Visceral pain: a review of experimental studies. Pain, 41, 167–234.CrossRefGoogle ScholarPubMed
Neuhuber, W. L. (1989). Vagal afferent fibers almost exclusively innervate islets in the rat pancreas as demonstrated by anterograde tracing. J. Auton. Nerv. Syst., 29, 13–18.CrossRefGoogle ScholarPubMed
Niijima, A. (1996). The afferent dischargee from sensors for interleukin 1 beta in the hepatoportal system in the anesthetized rat. J. Auton. Nerv. Syst., 61, 287–291.CrossRefGoogle Scholar
Nilsson, H., Jensen, P. E. and Mulvany, M. J. (1994). Minor role for direct adrenoceptor-mediated calcium entry in rat mesenteric small arteries. J. Vasc. Res., 31, 314–321.CrossRefGoogle ScholarPubMed
Nilsson, S. (1983). Autonomic Nerve Function in the Vertebrates. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Nilsson, S. and Holmgren, S. (eds.) (1994). The Autonomic Nervous System, Vol. 4, Comparative Physiology and Evolution of the Autonomic Nervous System, Chur, Switzerland: Harwood Academic Publishers.Google Scholar
Nisida, I. and Okada, H. (1960). The activity of the pupilloconstrictor centers. Jap. J. Physiol., 10, 64–72.CrossRefGoogle Scholar
Nja, A. and Purves, D. (1977a). Specific innervation of guinea-pig superior cervical ganglion cells by preganglionic fibres arising from different levels of the spinal cord. J. Physiol., 264, 565–583.CrossRefGoogle Scholar
Nja, A. and Purves, D. (1977b). Re-innervation of guinea-pig superior cervical ganglion cells by preganglionic fibres arising from different levels of the spinal cord. J. Physiol., 272, 633–651.CrossRefGoogle Scholar
Nolan, M. F., Logan, S. D. and Spanswick, D. (1999). Electrophysiological properties of electrical synapses between rat sympathetic preganglionic neurones in vitro. J. Physiol., 519, 753–764.CrossRefGoogle ScholarPubMed
Noll, G., Elam, M., Kunimoto, M., Karlsson, T. and Wallin, B. G. (1994). Skin sympathetic nerve activity and effector function during sleep in humans. Acta Physiol. Scand., 151, 319–329.CrossRefGoogle ScholarPubMed
Nordin, M. (1990). Sympathetic discharges in the human supraorbital nerve and their relation to sudo- and vasomotor responses. J. Physiol., 423, 241–255.CrossRefGoogle ScholarPubMed
Nordin, M. and Fagius, J. (1995). Effect of noxious stimulation on sympathetic vasoconstrictor outflow to human muscles. J. Physiol., 489, 885–894.CrossRefGoogle ScholarPubMed
Norman, R. A., Coleman, T. G.Jr and Dent, A. C. (1981). Continuous monitoring of arterial pressure indicates sinoaortic denervated rats are not hypertensive. Hypertension, 3, 119–125.CrossRefGoogle Scholar
North, R. A. (1986). Receptors on individual neurones. Neuroscience, 17, 899–907.CrossRefGoogle ScholarPubMed
Oberle, J., Elam, M., Karlsson, T. and Wallin, B. G. (1988). Temperature-dependent interaction between vasoconstrictor and vasodilator mechanisms in human skin. Acta Physiol. Scand., 132, 459–469.CrossRefGoogle ScholarPubMed
Obrist, P. A. (1981). Cardiovascular Psychophysiology: A Perspective. New York: Plenum Press.CrossRefGoogle Scholar
Ochoa, J. and Torebjörk, E. (1983). Sensations evoked by intraneural microstimulation of single mechanoreceptor units innervating the human hand. J. Physiol., 342, 633–654.CrossRefGoogle ScholarPubMed
Ochoa, J. and Torebjörk, H. E. (1989). Sensations evoked by intraneural microstimulation of C nociceptor fibres in human skin nerves. J. Physiol., 415, 583–599.CrossRefGoogle ScholarPubMed
Oh, E. J., Mazzone, S. B., Canning, B. J. and Weinreich, D. (2006). Reflex regulation of airway sympathetic nerves in guinea-pigs. J. Physiol., in press.CrossRefGoogle ScholarPubMed
Oldfield, B. J. and McLachlan, E. M. (1981). An analysis of the sympathetic preganglionic neurons projecting from the upper thoracic spinal roots of the cat. J. Comp. Neurol., 196, 329–345.CrossRefGoogle ScholarPubMed
Öngür, D. and Price, J. L. (2000). The organization of networks within the orbital and medial prefrontal cortex of rats, monkeys and humans. Cereb. Cortex, 10, 206–219.CrossRefGoogle ScholarPubMed
Öngür, D., An, X. and Price, J. L. (1998). Prefrontal cortical projections to the hypothalamus in macaque monkeys. J. Comp. Neurol., 401, 480–505.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Ootsuka, Y. and Terui, N. (1997). Functionally different neurons are organized topographically in the rostral ventrolateral medulla of rabbits. J. Auton. Nerv. Syst., 67, 67–78.CrossRefGoogle ScholarPubMed
Orr, R. and Marson, L. (1998). Identification of central nervous system neurons innervating the rat prostate: a transneuronal tracing study using pseudorabies virus. J. Auton. Nerv. Syst., 72, 4–15.CrossRefGoogle ScholarPubMed
Owens, N. C., Ootsuka, Y., Kanosue, K. and McAllen, R. M. (2002). Thermoregulatory control of sympathetic fibers supplying the rat's tail. J. Physiol., 543, 849–858.CrossRefGoogle ScholarPubMed
Owsjannikow, P. (1874). Über einen Unterschied in den reflectorischen Leistungen des verlängerten und des Rückenmarkes der Kaninchen [On a difference between reflexes mediated by the medulla oblongata and spinal cord in rabbits]. Sitzungsber. Akad. Wiss. Wien, Math.-Naturw., Abt. 2, 26, 457–464.Google Scholar
Ozaki, N. and Gebhart, G. F. (2001). Characterization of mechanosensitive splanchnic nerve afferent fibers innervating the rat stomach. Am. J. Physiol. Gastrointest. Liver Physiol., 281, G1449–G1459.CrossRefGoogle ScholarPubMed
Page, A. J. and Blackshaw, L. A. (1998). An in vitro study of the properties of vagal afferent fibres innervating the ferret oesophagus and stomach. J. Physiol., 512, 907–916.CrossRefGoogle Scholar
Page, A. J., Martin, C. M. and Blackshaw, L. A. (2002). Vagal mechanoreceptors and chemoreceptors in mouse stomach and esophagus. J. Neurophysiol., 87, 2095–2103.CrossRefGoogle ScholarPubMed
Paintal, A. S. (1973). Vagal sensory receptors and their reflex effects. Physiol. Rev., 53, 159–227.CrossRefGoogle ScholarPubMed
Paintal, A. S. (1986). The visceral sensations – some basic mechanisms. Prog. Brain Res., 67, 3–19.CrossRefGoogle ScholarPubMed
Pan, H. L. and Longhurst, J. C. (1996). Ischaemia-sensitive sympathetic afferents innervating the gastrointestinal tract function as nociceptors in cats. J. Physiol., 492, 841–850.CrossRefGoogle ScholarPubMed
Pan, H. L., Longhurst, J. C., Eisenach, J. C. and Chen, S. R. (1999). Role of protons in activation of cardiac sympathetic C-fibre afferents during ischaemia in cats. J. Physiol., 518, 857–866.CrossRefGoogle ScholarPubMed
Panksepp, J. (1998). Affective Neuroscience. New York, Oxford: Oxford University Press.Google Scholar
Papka, R. E., Williams, S., Miller, K. E., Copelin, T. and Puri, P. (1998). central nervous system location of uterine-related neurons revealed by trans-synaptic tracing with pseudorabies virus and their relation to estrogen receptor-immunoreactive neurons. Neuroscience, 84, 935–952.CrossRefGoogle ScholarPubMed
Parr, E. J. and Sharkey, K. A. (1996). Immunohistochemically-defined subtypes of neurons in the inferior mesenteric ganglion of the guinea-pig. J. Auton. Nerv. Syst., 59, 140–150.CrossRefGoogle ScholarPubMed
Paton, J. F. (1996a). The ventral medullary respiratory network of the mature mouse studied in a working heart–brainstem preparation. J. Physiol., 493, 819–831.CrossRefGoogle Scholar
Paton, J. F. (1996b). A working heart–brainstem preparation of the mouse. J. Neurosci. Methods, 65, 63–68.CrossRefGoogle Scholar
Paton, J. F. (1999). The Sharpey-Schafer prize lecture: nucleus tractus solitarii: integrating structures. Exp. Physiol., 84, 815–833.CrossRefGoogle ScholarPubMed
Paton, J. F. and Dutschmann, M. (2002). Central control of upper airway resistance regulating respiratory airflow in mammals. J. Anat., 201, 319–323.CrossRefGoogle ScholarPubMed
Paton, J. F. and Kasparov, S. (2000). Sensory channel specific modulation in the nucleus of the solitary tract. J. Auton. Nerv. Syst., 80, 117–129.CrossRefGoogle ScholarPubMed
Paton, J. F., Li, Y. W., Deuchars, J. and Kasparov, S. (2000). Properties of solitary tract neurons receiving inputs from the sub-diaphragmatic vagus nerve. Neuroscience, 95, 141–153.CrossRefGoogle ScholarPubMed
Paton, J. F., Deuchars, J., Li, Y. W. and Kasparov, S. (2001). Properties of solitary tract neurones responding to peripheral arterial chemoreceptors. Neuroscience, 105, 231–248.CrossRefGoogle ScholarPubMed
Paton, J. F., Deuchar, J., Wang, S. and Kasparov, S. (2005). Nitroxergic modulation in the nucleus tractus solitarii: implications for cardiovascular function. In Advances in Vagal Afferent Neurobiology, ed. Undem, B. and Weinreich, D.. Boca Raton: CRC Press, pp. 209–246.CrossRefGoogle Scholar
Paxinos, G. and Watson, C. (eds.) (1998). The Rat Brain. San Diego: Academic Press.Google Scholar
Peng, H., Matchkov, V., Ivarsen, A., Aalkjaer, C. and Nilsson, H. (2001). Hypothesis for the initiation of vasomotion. Circ. Res., 88, 810–815.CrossRefGoogle ScholarPubMed
Pernow, J. (1988). Co-release and functional interactions of neuropeptide Y and noradrenaline in peripheral sympathetic vascular control. Acta Physiol. Scand. Suppl., 568, 1–56.Google ScholarPubMed
Perry, M. J. and Lawson, S. N. (1998). Differences in expression of oligosaccharides, neuropeptides, carbonic anhydrase and neurofilament in rat primary afferent neurons retrogradely labelled via skin, muscle or visceral nerves. Neuroscience, 85, 293–310.CrossRefGoogle ScholarPubMed
Persson, P., Ehmke, H., Kirchheim, H. and Seller, H. (1988). Effect of sino-aortic denervation in comparison to cardiopulmonary deafferentiation on long-term blood pressure in conscious dogs. Pflügers Arch., 411, 160–166.CrossRefGoogle ScholarPubMed
Petras, J. M. and Cummings, J. F. (1972). Autonomic neurons in the spinal cord of the Rhesus monkey: a correlation of the findings of cytoarchitectonics and sympathectomy with fiber degeneration following dorsal rhizotomy. J. Comp. Neurol., 146, 189–218.CrossRefGoogle ScholarPubMed
Petrovich, G. D., Canteras, N. S. and Swanson, L. W. (2001). Combinatorial amygdalar inputs to hippocampal domains and hypothalamic behavior systems. Brain Res. Brain Res. Rev., 38, 247–289.CrossRefGoogle ScholarPubMed
Petty, B. G., Cornblath, D. R., Adornato, B. T.et al. (1994). The effect of systemically administered recombinant human nerve growth factor in healthy human subjects. Ann. Neurol., 36, 244–246.CrossRefGoogle ScholarPubMed
Phan, K. L., Wager, T., Taylor, S. F. and Liberzon, I. (2002). Functional neuroanatomy of emotion: a meta-analysis of emotion activation studies in PET and fMRI. Neuroimage, 16, 331–348.CrossRefGoogle ScholarPubMed
Phillips, J. K., Goodchild, A. K., Dubey, R.et al. (2001). Differential expression of catecholamine biosynthetic enzymes in the rat ventrolateral medulla. J. Comp. Neurol., 432, 20–34.CrossRefGoogle ScholarPubMed
Phillips, R. J. and Powley, T. L. (2000). Tension and stretch receptors in gastrointestinal smooth muscle: re-evaluating vagal mechanoreceptor electrophysiology. Brain Res. Brain Res. Rev., 34, 1–26.CrossRefGoogle ScholarPubMed
Phillips, R. J., Baronowsky, E. A. and Powley, T. L. (1997). Afferent innervation of gastrointestinal tract smooth muscle by the hepatic branch of the vagus. J. Comp. Neurol., 384, 248–270.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Pick, J. (1970). The Autonomic Nervous System. Philadelphia: Lippincott.Google Scholar
Pickering, A. E., Boscan, P. and Paton, J. F. (2003). Nociception attenuates parasympathetic but not sympathetic baroreflex via neurokinin 1 receptors in the rat nucleus tractus solitarii. J. Physiol., 551, 589–599.CrossRefGoogle Scholar
Pierce, P. A., Xie, G. X., Peroutka, S. J., Green, P. G. and Levine, J. D. (1995). 5-Hydroxytryptamine-induced synovial plasma extravasation is mediated via 5-hydroxytryptamine2 A receptors on sympathetic efferent terminals. J. Pharmacol. Exp. Ther., 275, 502–508.Google Scholar
Pilowsky, P. M. and Goodchild, A. K. (2002). Baroreceptor reflex pathways and neurotransmitters: 10 years on. J. Hypertens., 20, 1675–1688.CrossRefGoogle Scholar
Pilowsky, P. M. and Makeham, J. (2001). Juxtacellular labeling of identified neurons: kiss the cells and make them dye. J. Comp. Neurol., 433, 1–3.CrossRefGoogle ScholarPubMed
Pilowsky, P. M., Jiang, C. and Lipski, J. (1990). An intracellular study of respiratory neurons in the rostral ventrolateral medulla of the rat and their relationship to catecholamine-containing neurons. J. Comp. Neurol., 301, 604–617.CrossRefGoogle ScholarPubMed
Pilowsky, P., Llewellyn-Smith, I. J., Arnolda, L., Minson, J. and Chalmers, J. (1994). Intracellular recording from sympathetic preganglionic neurons in cat lumbar spinal cord. Brain Res., 656, 319–328.CrossRefGoogle ScholarPubMed
Pinault, D. (1996). A novel single-cell staining procedure performed in vivo under electrophysiological control: morpho-functional features of juxtacellularly labeled thalamic cells and other central neurons with biocytin or Neurobiotin. J. Neurosci. Methods, 65, 113–136.CrossRefGoogle ScholarPubMed
Polgar, E., Gray, S., Riddell, J. S. and Todd, A. J. (2004). Lack of evidence for significant neuronal loss in laminae I-III of the spinal dorsal horn of the rat in the chronic constriction injury model. Pain, 111, 144–150.CrossRefGoogle ScholarPubMed
Polson, J. W., Halliday, G. M., McAllen, R. M., Coleman, M. J. and Dampney, R. A. (1992). Rostrocaudal differences in morphology and neurotransmitter content of cells in the subretrofacial vasomotor nucleus. J. Auton. Nerv. Syst., 38, 117–138.CrossRefGoogle ScholarPubMed
Polson, J. W., Potts, P. D., Li, Y. W. and Dampney, R. A. (1995). Fos expression in neurons projecting to the pressor region in the rostral ventrolateral medulla after sustained hypertension in conscious rabbits. Neuroscience, 67, 107–123.CrossRefGoogle ScholarPubMed
Popescu, L. M., Ciontea, S. M., Cretoiu, D.et al. (2005a). Novel type of interstitial cell (Cajal-like) in human fallopian tube. J. Cell Mol. Med., 9, 479–523.CrossRefGoogle Scholar
Popescu, L. M., Hinescu, M. E., Ionescu, N.et al. (2005b). Interstitial cells of Cajal in pancreas. J. Cell Mol. Med., 9, 169–190.CrossRefGoogle Scholar
Possas, O. S., Campos, R. R. Jr., Cravo, S. L., Lopes, O. U. and Guertzenstein, P. G. (1994). A fall in arterial blood pressure produced by inhibition of the caudalmost ventrolateral medulla: the caudal pressor area. J. Auton. Nerv. Syst., 49, 235–245.CrossRefGoogle ScholarPubMed
Potter, E. K. (1987). Guanethidine blocks neuropeptide-Y-like inhibitory action of sympathetic nerves on cardiac vagus. J. Auton. Nerv. Syst., 21, 87–90.CrossRefGoogle ScholarPubMed
Potter, E. K. (1991). Neuropeptide Y as an autonomic neurotransmitter. In Novel Peripheral Neurotransmitter, ed. Bell, C.. New York: Pergamon, pp. 81–112.Google Scholar
Powley, T. L., Berthoud, H. R., Fox, A. P. and Laughton, W. (1992). The dorsal vagal complex forms a sensory-motor lattice: the circuitry of gastrointestinal reflexes. In Neuroanatomy and Physiology of Abdominal Vagal Afferents, ed. Ritter, S., Ritter, R. C. and Barnes, C. D.. Boca Raton: CRC Press, pp. 55–79.Google Scholar
Prechtl, J. C. and Powley, T. L. (1985). Organization and distribution of the rat subdiaphragmatic vagus and associated paraganglia. J. Comp. Neurol., 235, 182–195.CrossRefGoogle ScholarPubMed
Prechtl, J. C. and Powley, T. L. (1990a). The fiber composition of the abdominal vagus of the rat. Anat. Embryol. (Berl.), 181, 101–115.CrossRefGoogle Scholar
Prechtl, J. C. and Powley, T. L. (1990b). B-afferents: a fundamental division of the nervous system mediating homeostasis [the article includes Open Peer Commentary]. Behav. Brain Sci., 13, 289–331.CrossRefGoogle Scholar
Purves, D. (1988). Body and Brain: A Tropic Theory of Neural Connection. Cambridge: Harvard University Press.Google Scholar
Purves, D. and Hume, R. I. (1981). The relation of postsynaptic geometry to the number of presynaptic axons that innervate autonomic ganglion cells. J. Neurosci., 1, 441–452.CrossRefGoogle ScholarPubMed
Purves, D. and Lichtman, J. W. (1978). Formation and maintenance of synaptic connections in autonomic ganglia. Physiol. Rev., 58, 821–862.CrossRefGoogle ScholarPubMed
Purves, D. and Lichtman, J. W. (1985). Geometrical differences among homologous neurons in mammals. Science, 228, 298–302.CrossRefGoogle ScholarPubMed
Purves, D. and Wigston, D. J. (1983). Neural units in the superior cervical ganglion of the guinea-pig. J. Physiol., 334, 169–178.CrossRefGoogle ScholarPubMed
Purves, D., Rubin, E., Snider, W. D. and Lichtman, J. (1986). Relation of animal size to convergence, divergence, and neuronal number in peripheral sympathetic pathways. J. Neurosci., 6, 158–163.CrossRefGoogle ScholarPubMed
Purves, D., Snider, W. D. and Voyvodic, J. T. (1988). Trophic regulation of nerve cell morphology and innervation in the autonomic nervous system. Nature, 336, 123–128.CrossRefGoogle ScholarPubMed
Pyner, S. and Coote, J. H. (1994). Evidence that sympathetic preganglionic neurons are arranged in target-specific columns in the thoracic spinal cord of the rat. J. Comp. Neurol., 342, 15–22.CrossRefGoogle ScholarPubMed
Pyner, S. and Coote, J. H. (1998). Rostroventrolateral medulla neurons preferentially project to target-specified sympathetic preganglionic neurons. Neuroscience, 83, 617–631.CrossRefGoogle ScholarPubMed
Qin, C., Chandler, M. J., Miller, K. E. and Foreman, R. D. (2001). Responses and afferent pathways of superficial and deeper c(1)–c(2) spinal cells to intrapericardial algogenic chemicals in rats. J. Neurophysiol., 85, 1522–1532.CrossRefGoogle ScholarPubMed
Quigg, M., Elfvin, L. G. and Aldskogius, H. (1990). Anterograde transsynaptic transport of WGA-horse radish peroxidase from spinal afferents to postganglionic sympathetic cells of the stellate ganglion of the guinea pig. Brain Res., 518, 173–178.CrossRefGoogle Scholar
Randall, W. C. (ed.) (1984). Nervous Control of Cardiovascular Function. New York, Oxford: Oxford University Press.Google Scholar
Randall, W. C., Wurster, R. D. and Lewin, R. J. (1966). Responses of patients with high spinal transection to high ambient temperatures. J. Appl. Physiol., 21, 985–993.CrossRefGoogle ScholarPubMed
Randich, A. and Gebhart, G. F. (1992). Vagal afferent modulation of nociception. Brain Res. Rev., 17, 77–99.CrossRefGoogle ScholarPubMed
Ranson, S. W. and Clark, S. L. (1959). The Anatomy of the Nervous System. 10th edn. Philadelphia, London: W. B. Saunders Company.Google Scholar
Ranson, S. W. and Magoun, H. W. (1939). The hypothalamus. Ergebn. Physiol., 41, 56–163.CrossRefGoogle Scholar
Rathner, J. A. and McAllen, R. M. (1999). Differential control of sympathetic drive to the rat tail artery and kidney by medullary premotor cell groups. Brain Res., 834, 196–199.CrossRefGoogle ScholarPubMed
Rathner, J. A., Owens, N. C. and McAllen, R. M. (2001). Cold-activated raphe-spinal neurons in rats. J. Physiol., 535, 841–854.CrossRefGoogle ScholarPubMed
Reed, D. E. and Vanner, S. J. (2003). Long vasodilator reflexes projecting through the myenteric plexus in guinea-pig ileum. J. Physiol., 553, 911–924.CrossRefGoogle ScholarPubMed
Reiner, A., Karten, H. J., Gamlin, P. D. R. and Erichsen, J. T. (1983). Parasympathetic ocular control. Functional subdivisions and circuitry of the avian nucleus Edinger–Westphal. Trends Neurosci., 6, 140–145.CrossRefGoogle Scholar
Reis, D. J., Golanov, E. V., Ruggiero, D. A. and Sun, M. K. (1994). Sympatho-excitatory neurons of the rostral ventrolateral medulla are oxygen sensors and essential elements in the tonic and reflex control of the systemic and cerebral circulations. J. Hypertens. Suppl., 12, S159–S180.Google ScholarPubMed
Reis, D. J., Golanov, E. V., Galea, E. and Feinstein, D. L. (1997). Central neurogenic neuroprotection: central neural systems that protect the brain from hypoxia and ischemia. Ann. New York Acad. Sci., 835, 168–186.CrossRefGoogle ScholarPubMed
Rekling, J. C. and Feldman, J. L. (1998). PreBotzinger complex and pacemaker neurons: hypothesized site and kernel for respiratory rhythm generation. Annu. Rev. Physiol., 60, 385–405.CrossRefGoogle ScholarPubMed
Rexed, B. (1952). The cytoarchitectonic organization of the spinal cord in the cat. J. Comp. Neurol., 96, 414–495.CrossRefGoogle ScholarPubMed
Rexed, B. (1954). A cytoarchitectonic atlas of the spinal cord in the cat. J. Comp. Neurol., 100, 297–379.CrossRefGoogle ScholarPubMed
Richards, W., Hillsley, K., Eastwood, C. and Grundy, D. (1996). Sensitivity of vagal mucosal afferents to cholecystokinin and its role in afferent signal transduction in the rat. J. Physiol., 497, 473–481.CrossRefGoogle ScholarPubMed
Richter, D. W. (1982). Generation and maintenance of the respiratory rhythm. J. Exp. Biol., 100, 93–107.Google ScholarPubMed
Richter, D. W. (1996). Neural regulation of respiration: rhythmogenesis and afferent control. In Comprehensive Human Physiology, ed. Greger, R. and Windhorst, U.. Berlin, Heidelberg, New York: Springer, pp. 2079–2095.CrossRefGoogle Scholar
Richter, D. W. and Ballantyne, D. (1983). A three phase theory about the basic respiratory pattern generator. In Central Neurone Environment, ed. Schläfke, M. E., Koepchen, H. P. and See, W. R.. Berlin, Heidelberg, New York: Springer, pp. 164–174.Google Scholar
Richter, D. W. and Spyer, K. M. (1990). Cardiorespiratory control. In Central Regulation of Autonomic Functions, eds. Loewy, A. D. and Spyer, K. M.. New York, Oxford: Oxford University Press, pp. 189–207.Google Scholar
Richter, D. W. and Spyer, K. M. (2001). Studying rhythmogenesis of breathing: comparison of in vivo and in vitro models. Trends Neurosci., 24, 464–472.CrossRefGoogle ScholarPubMed
Richter, D. W., Spyer, K. M., Gilbey, M. P. et al. (1991). On the existence of a common cardiorespiratory network. In Cardiorespiratory and Motor Coordination, ed. Koepchen, H. P. and Huopaniemi, T.. Berlin, Heidelberg, New York: Springer, pp. 118–130.CrossRefGoogle Scholar
Rinaman, L., Card, J. P., Schwaber, J. S. and Miselis, R. R. (1989). Ultrastructural demonstration of a gastric monosynaptic vagal circuit in the nucleus of the solitary tract in rat. J. Neurosci., 9, 1985–1996.CrossRefGoogle ScholarPubMed
Risold, P. Y., Thompson, R. H. and Swanson, L. W. (1997). The structural organization of connections between hypothalamus and cerebral cortex. Brain Res. Brain Res. Rev., 24, 197–254.CrossRefGoogle ScholarPubMed
Ritter, S., Ritter, R. C. and Barnes, C. D. (eds.) (1992). Neuroanatomy and Physiology of Abdominal Vagal Afferents. Boca Raton: CRC Press.Google Scholar
Robertson, D. and Biaggioni, I. (eds.) (1995). The Autonomic Nervous System, Vol. 5, Disorders of the Autonomic Nervous System. Luxembourg: Harwood Academic Publishers.
Roddie, I. C. (1977). Human responses to emotional stress. Irish J. Med. Sci., 146, 395–417.CrossRefGoogle ScholarPubMed
Roddie, I. C., Shepherd, J. T. and Whelan, R. F. (1957). The contribution of constrictor and dilator nerves to the skin vasodilation during body heating. J. Physiol., 136, 489–497.CrossRefGoogle ScholarPubMed
Rogers, R. C. and Hermann, G. E. (1992). Central regulation of brainstem gastric vago-vagal control circuits. In Neuroanatomy and Physiology of Abdominal Vagal Afferents, ed. Ritter, S., Ritter, R. C. and Barnes, C. D.. Boca Raton: CRC Press, pp. 99–134.Google Scholar
Rogers, R. C., Kita, H., Butcher, L. L. and Novin, D. (1980). Afferent projections to the dorsal motor nucleus of the vagus. Brain Res. Bull., 5, 365–373.CrossRefGoogle ScholarPubMed
Rogers, R. C., McTigue, D. M. and Hermann, G. E. (1995). Vagovagal reflex control of digestion: afferent modulation by neural and “endoneurocrine” factors. Am. J. Physiol., 268, G1–G10.Google ScholarPubMed
Rogers, R. C., Hermann, G. E. and Travagli, R. A. (1999). Brainstem pathways responsible for oesophageal control of gastric motility and tone in the rat. J. Physiol., 514, 369–383.CrossRefGoogle ScholarPubMed
Rogers, R. C., Travagli, R. A. and Hermann, G. E. (2003). Noradrenergic neurons in the rat solitary nucleus participate in the esophageal-gastric relaxation reflex. Am. J. Physiol. Regul. Integr. Comp. Physiol., 285, R479–R489.CrossRefGoogle ScholarPubMed
Rogers, R. F., Paton, J. F. and Schwaber, J. S. (1993). nucleus tractus solitarii neuronal responses to arterial pressure and pressure changes in the rat. Am. J. Physiol., 265, R1355–R1368.Google ScholarPubMed
Roman, C. and Gonella, J. (1994). Extrinsic control of digestive tract motility. In Physiology of the Gastrointestinal Tract, 3rd edn., ed. , L. R. Johnson. New York: Raven Press, pp. 507–553.Google Scholar
Romano, T. A., Felten, S. Y., Felten, D. L. and Olschowka, J. A. (1991). Neuropeptide-Y innervation of the rat spleen: another potential immunomodulatory neuropeptide. Brain Behav. Immun., 5, 116–131.CrossRefGoogle ScholarPubMed
Root, W. S. and Bard, P. (1947). The mediation of feline erection through sympathetic pathways with some remarks on sexual behavior after deafferentation of the genitalia. Am. J. Physiol., 151, 80–90.Google ScholarPubMed
Rosell, S. (1980). Neuronal control of microvessels. Annu. Rev. Physiol., 42, 359–371.CrossRefGoogle ScholarPubMed
Rosell, S. and Belfrage, E. (1979). Blood circulation in adipose tissue. Physiol. Rev., 59, 1078–1104.CrossRefGoogle ScholarPubMed
Rosicka, M., Krsek, M., Jarkovska, Z., Marek, J. and Schreiber, V. (2002). Ghrelin – a new endogenous growth hormone secretagogue. Physiol. Res., 51, 435–441.Google ScholarPubMed
Ross, C. A., Ruggiero, D. A., Park, D. H.et al. (1984b). Tonic vasomotor control by the rostral ventrolateral medulla: effect of electrical or chemical stimulation of the area containing C1 adrenaline neurons on arterial pressure, heart rate, and plasma catecholamines and vasopressin. J. Neurosci., 4, 474–494.CrossRefGoogle Scholar
Ross, C. A., Ruggiero, D. A., Joh, T. H., Park, D. H. and Reis, D. J. (1984a). Rostral ventrolateral medulla: selective projections to the thoracic autonomic cell column from the region containing C1 adrenaline neurons. J. Comp. Neurol., 228, 168–185.CrossRefGoogle Scholar
Rossiter, C. D., Norman, W. P., Jain, M.et al. (1990). Control of lower esophageal sphincter pressure by two sites in dorsal motor nucleus of the vagus. Am. J. Physiol., 259, G899–G906.Google ScholarPubMed
Rowell, L. B. (1993). Human Cardiovascular Control. New York, Oxford: Oxford University Press.Google Scholar
Ruggiero, D. A., Cravo, S. L., Arango, V. and Reis, D. J. (1989). Central control of the circulation by the rostral ventrolateral reticular nucleus: anatomical substrates. Prog. Brain Res., 81, 49–79.CrossRefGoogle ScholarPubMed
Rushmer, R. F. and Smith, O. A. Jr. (1959). Cardiac control. Physiol. Rev., 39, 41–68.CrossRefGoogle ScholarPubMed
Rybak, I. A., Paton, J. F. R., Rogers, R. F. and St.-John, W. M. (2002). Generation of the respiratory rhythm: state-dependency and switching. Neurocomputing, 44–46, 605–614.CrossRefGoogle Scholar
Rybak, I. A., Shevtsova, N. A., St.-John, W. M., Paton, J. F. and Pierrefiche, O. (2003). Endogenous rhythm generation in the pre-Bötzinger complex and ionic currents: modelling and in vitro studies. Eur. J. Neurosci., 18, 239–257.CrossRefGoogle ScholarPubMed
Rybak, I. A., Shevtsova, N. A., Paton, J. F.et al. (2004). Modeling the ponto-medullary respiratory network. Respir. Physiol. Neurobiol., 143, 307–319.CrossRefGoogle ScholarPubMed
Sanders, K. M. and Smith, T. K. (2003). Neural regulation of colonic motor function. In Textbook of Colonic Disease, ed. Koch, T.. Totowa, New Jersey, USA: Humana Press, Inc., pp. 35–52.CrossRefGoogle Scholar
Sanders, K. M., Ordog, T., Koh, S. D., Torihashi, S. and Ward, S. M. (1999). Development and plasticity of interstitial cells of Cajal. Neurogastroenterol. Motil., 11, 311–338.CrossRefGoogle ScholarPubMed
Sanders, K. M., Ordog, T., Koh, S. D. and Ward, S. M. (2000). A novel pacemaker mechanism drives gastrointestinal rhythmicity. News Physiol. Sci., 15, 291–298.Google ScholarPubMed
Santicioli, P. and Maggi, C. A. (1998). Myogenic and neurogenic factors in the control of pyeloureteral motility and ureteral peristalsis. Pharmacol. Rev., 50, 683–722.Google ScholarPubMed
Saper, C. B. (1995). Central autonomic system. In The Rat Nervous System, ed. Paxinos, G.. San Diego: Academic Press, pp. 107–135.Google Scholar
Saper, C. B. (2002). The central autonomic nervous system: conscious visceral perception and autonomic pattern generation. Annu. Rev. Neurosci., 25, 433–469.CrossRefGoogle ScholarPubMed
Saper, C. B. (2004). Anatomy of hypothalamus. In The Human Nervous System, ed. Paxinos, G. and Mai, J. K.. Amsterdam: Elsevier Academic Press, pp. 513–550.Google Scholar
Saphier, D. (1993). Psychoimmunology: the missing link. In Hormonally Induced Changes in Mind and Brain, ed. Schulkin, J.. Boston, New York: Academic Press, pp. 191–224.Google Scholar
Sato, A. (1972). Somato-sympathetic reflex discharges evoked through supramedullary pathways. Pflügers Arch., 332, 117–126.CrossRefGoogle ScholarPubMed
Sato, A. and Schmidt, R. F. (1971). Spinal and supraspinal components of the reflex discharges into lumbar and thoracic white rami. J. Physiol., 212, 839–850.CrossRefGoogle ScholarPubMed
Sato, A. and Schmidt, R. F. (1973). Somatosympathetic reflexes: afferent fibers, central pathways, discharge characteristics. Physiol. Rev., 53, 916–947.CrossRefGoogle ScholarPubMed
Sato, A., Sato, Y. and Schmidt, R. F. (1997). The impact of somatosensory input on autonomic functions. Rev. Physiol. Biochem. Pharmacol., 130, 1–328.CrossRefGoogle ScholarPubMed
Schaible, H. G. and Schmidt, R. F. (1988). Time course of mechanosensitivity changes in articular afferents during a developing experimental arthritis. J. Neurophysiol., 60, 2180–2195.CrossRefGoogle ScholarPubMed
Schelegle, E. S. and Green, J. F. (2001). An overview of the anatomy and physiology of slowly adapting pulmonary stretch receptors. Respir. Physiol., 125, 17–31.CrossRefGoogle ScholarPubMed
Schläfke, M. E. and Loeschke, H. H. (1967). Lokalisation eines an der Regulation von Atmung und Kreislauf beteiligten Gebietes an der ventralen Oberfläche der Medulla oblongata durch Kälteblockade [Localization of a center on the ventral surface of the medulla oblongata involved in regulation of respiratory and cardiovascular system by cold blockade]. Pflügers Arch., 297, 201–220.CrossRefGoogle Scholar
Schmelz, M., Michael, K., Weidner, C.et al. (2000). Which nerve fibers mediate the axon reflex flare in human skin?NeuroReport, 11, 645–648.CrossRefGoogle ScholarPubMed
Schmidt, R., Schmelz, M., Forster, C.et al. (1995). Novel classes of responsive and unresponsive C nociceptors in human skin. J. Neurosci., 15, 333–341.CrossRefGoogle ScholarPubMed
Schmidt, R., Schmelz, M., Torebjörk, H. E. and Handwerker, H. O. (2000). Mechano-insensitive nociceptors encode pain evoked by tonic pressure to human skin. Neuroscience, 98, 793–800.CrossRefGoogle ScholarPubMed
Schmidt, R., Schmelz, M., Weidner, C., Handwerker, H. O. and Torebjörk, H. E. (2002). Innervation territories of mechano-insensitive C nociceptors in human skin. J. Neurophysiol., 88, 1859–1866.CrossRefGoogle ScholarPubMed
Schmidt-Vanderheyden, W. and Koepchen, H. P. (1967). Zum Mechanismus der Adrenalindilatation der Skeletmuskelgefäße [On the mechanism of adrenaline dilation of skeletal muscle blood vessels]. Pflügers Arch., 298, 1–11.CrossRefGoogle Scholar
Schomburg, E. D. (1990). Spinal sensorimotor systems and their supraspinal control. Neurosci. Res., 7, 265–340.CrossRefGoogle ScholarPubMed
Schramm, L. P., Strack, A. M., Platt, K. B. and Loewy, A. D. (1993). Peripheral and central pathways regulating the kidney: a study using pseudorabies virus. Brain Res., 616, 251–262.CrossRefGoogle ScholarPubMed
Schreihofer, A. M. and Guyenet, P. G. (1997). Identification of C1 presympathetic neurons in rat rostral ventrolateral medulla by juxtacellular labeling in vivo. J. Comp. Neurol., 387, 524–536.3.0.CO;2-4>CrossRefGoogle ScholarPubMed
Schreihofer, A. M. and Guyenet, P. G. (2000). Sympathetic reflexes after depletion of bulbospinal catecholaminergic neurons with anti-dopamine-β-hydroxylase-saporin. Am. J. Physiol. Regul. Integr. Comp. Physiol., 279, R729–R742.CrossRefGoogle Scholar
Schreihofer, A. M. and Guyenet, P. G. (2002). The baroreflex and beyond: control of sympathetic vasomotor tone by γ-aminobutyric acidergic neurons in the ventrolateral medulla. Clin. Exp. Pharmacol. Physiol., 29, 514–521.CrossRefGoogle Scholar
Schreihofer, A. M. and Guyenet, P. G. (2003). Baro-activated neurons with pulse-modulated activity in the rat caudal ventrolateral medulla express GAD67 messenger ribonucleic acid. J. Neurophysiol., 89, 1265–1277.CrossRefGoogle Scholar
Schreihofer, A. M., Stornetta, R. L. and Guyenet, P. G. (1999). Evidence for glycinergic respiratory neurons: Bötzinger neurons express messenger ribonucleic acid for glycinergic transporter 2. J. Comp. Neurol., 407, 583–597.3.0.CO;2-E>CrossRefGoogle ScholarPubMed
Schreihofer, A. M., Stornetta, R. L. and Guyenet, P. G. (2000). Regulation of sympathetic tone and arterial pressure by rostral ventrolateral medulla after depletion of C1 cells in rat. J. Physiol., 529, 221–236.CrossRefGoogle ScholarPubMed
Schulkin, J. (2003a). Allostasis: a neural behavioral perspective. Horm. Behav., 43, 21–27.CrossRefGoogle Scholar
Schulkin, J. (2003b). Rethinking Homeostasis. Allostatic Regulation in Physiology and Pathophysiology. Cambridge Massachusetts: The MIT Press.Google Scholar
Schwartz, M. W. and Morton, G. J. (2002). Obesity: keeping hunger at bay. Nature, 418, 595–597.CrossRefGoogle ScholarPubMed
Schwarzacher, S. W., Wilhelm, Z., Anders, K. and Richter, D. W. (1991). The medullary respiratory network in the rat. J. Physiol., 435, 631–644.CrossRefGoogle ScholarPubMed
Seagard, J. L., Gallenberg, L. A., Hopp, F. A. and Dean, C. (1992). Acute resetting in two functionally different types of carotid baroreceptors. Circ. Res., 70, 559–565.CrossRefGoogle ScholarPubMed
Seals, D. R., Suwarno, N. O. and Dempsey, J. A. (1990). Influence of lung volume on sympathetic nerve discharge in normal humans. Circ. Res., 67, 130–141.CrossRefGoogle ScholarPubMed
Seals, D. R., Suwarno, N. O., Joyner, M. J.et al. (1993). Respiratory modulation of muscle sympathetic nerve activity in intact and lung denervated humans. Circ. Res., 72, 440–454.CrossRefGoogle ScholarPubMed
Seller, H., Langhorst, P., Richter, D. and Koepchen, H. P. (1968). Über die Abhängigkeit der pressoreceptorischen Hemmung des Sympathicus von der Atemphase und ihre Auswirkung in der Vasomotorik [Respiratory variations of baroreceptor reflex transmission and their effects on sympathetic activity and vasomotor tone]. Pflügers Arch., 302, 300–314.CrossRefGoogle Scholar
Selyanko, A. A. (1992). Membrane properties and firing characteristics of rat cardiac neurones in vitro. J. Auton. Nerv. Syst., 39, 181–189.CrossRefGoogle ScholarPubMed
Semans, J. H. and Langworthy, O. R. (1938). Observations on the neurophysiology of sexual function in the male cat. J. Urol., 40, 836–846.CrossRefGoogle Scholar
Sengupta, J. N. and Gebhart, G. F. (1994a). Characterization of mechanosensitive pelvic nerve afferent fibers innervating the colon of the rat. J. Neurophysiol., 71, 2046–2060.CrossRefGoogle Scholar
Sengupta, J. N. and Gebhart, G. F. (1994b). Mechanosensitive properties of pelvic nerve afferent fibers innervating the urinary bladder of the rat. J. Neurophysiol., 72, 2420–2430.CrossRefGoogle Scholar
Sengupta, J. N. and Gebhart, G. F. (1995). Mechanosensitive afferent fibers in the gastrointestinal and lower urinary tracts. In Visceral Pain, ed. Gebhart, G. F.. Seattle: IASP Press, pp. 75–98.Google Scholar
Sengupta, J. N., Saha, J. K. and Goyal, R. K. (1990). Stimulus–response function studies of esophageal mechanosensitive nociceptors in sympathetic afferents of opossum. J. Neurophysiol., 64, 796–812.CrossRefGoogle ScholarPubMed
Sengupta, J. N., Su, X. and Gebhart, G. F. (1996). Kappa, but not mu or delta, opioids attenuate responses to distention of afferent fibers innervating the rat colon. Gastroenterology, 111, 968–980.CrossRefGoogle ScholarPubMed
Shade, R. E., Bishop, V. S., Haywood, J. R. and Hamm, C. K. (1990). Cardiovascular and neuroendocrine responses to baroreceptor denervation in baboons. Am. J. Physiol., 258, R930–R938.Google ScholarPubMed
Shafton, A. D., Oldfield, B. J. and McAllen, R. M. (1992). CRF-like immunoreactivity selectively labels preganglionic sudomotor neurons in cat. Brain Res., 599, 253–260.CrossRefGoogle ScholarPubMed
Shah, S. D., Tse, T. F., Clutter, W. E. and Cryer, P. E. (1984). The human sympathochromaffin system. Am. J. Physiol., 247, E380–E384.Google ScholarPubMed
Shanahan, F. (1994). The intestinal immune system. In Physiology of the Gastrointestinal Tract, 3rd edn., ed. Johnson, L. R.. New York: Raven Press, pp. 643–684.Google Scholar
Sharkey, K. A. and Mawe, G. M. (2002). Neuroimmune and epithelial interactions in intestinal inflammation. Curr. Opin. Pharmacol., 2, 669–677.CrossRefGoogle ScholarPubMed
Sheehan, D. (1936). Discovery of the autonomic nervous system. Arch. Neurol. Psychiat., 35, 1081–1115.CrossRefGoogle Scholar
Sheehan, D. (1941). The autonomic nervous system prior to Gaskell. New Engl. J. Med., 224, 457–460.CrossRefGoogle Scholar
Shefchyk, S. J. (2001). Sacral spinal interneurons and the control of urinary bladder and urethral striated sphincter muscle function. J. Physiol., 533, 57–63.CrossRefGoogle ScholarPubMed
Shefchyk, S. J. (2002). Spinal cord neural organization controlling the urinary bladder and striated sphincter. Prog. Brain Res., 137, 71–82.CrossRefGoogle ScholarPubMed
Shepherd, J. Z. and Vatner, S. F. (eds.) (1996). Nervous Control of the Heart. Vol. 9, The Autonomic Nervous System, Amsterdam: Harwood Academic Publishers.Google Scholar
Sherbourne, C. D., Gonzales, R., Goldyne, M. E. and Levine, J. D. (1992). Norepinephrine-induced increase in sympathetic neuron-derived prostaglandins is independent of neuronal release mechanisms. Neurosci. Lett., 139, 188–190.CrossRefGoogle ScholarPubMed
Sherrington, C. S. (1900). Cutaneous sensation. In Textbook of Physiology, Vol. 2, ed. Schäfer, E. A.. Edinburgh, London: Young J. Pentland, pp. 920–1001.Google Scholar
Sherrington, C. (1947) [1906]. The Integrative Action of the Nervous System, 2nd edn. New Haven: Yale University Press.Google Scholar
Sillito, A. M. and Zbrozyna, A. W. (1970a). The localization of pupilloconstrictor function within the mid-brain of the cat. J. Physiol., 211, 461–477.CrossRefGoogle Scholar
Sillito, A. M. and Zbrozyna, A. W. (1970b). The activity characteristics of the preganglionic pupilloconstrictor neurones. J. Physiol., 211, 767–779.CrossRefGoogle Scholar
Silva-Carvalho, L., Paton, J. F., Goldsmith, G. E. and Spyer, K. M. (1991). The effects of electrical stimulation of lobule IXb of the posterior cerebellar vermis on neurones within the rostral ventrolateral medulla in the anaesthetised cat. J. Auton. Nerv. Syst., 36, 97–106.CrossRefGoogle ScholarPubMed
Silverberg, A. B., Shah, S. D., Haymond, M. W. and Cryer, P. E. (1978). Norepinephrine: hormone and neurotransmitter. Am. J. Physiol., 234, E252–E256.Google ScholarPubMed
Simon, E. (1974). Temperature regulation: the spinal cord as a site of extrahypothalamic thermoregulatory functions. Rev. Physiol. Biochem. Pharmacol., 71, 1–76.CrossRefGoogle Scholar
Simon, E., Pierau, F. K. and Taylor, D. C. (1986). Central and peripheral thermal control of effectors in homeothermic temperature regulation. Physiol. Rev., 66, 235–300.CrossRefGoogle ScholarPubMed
Simone, D. A., Sorkin, L. S., Oh, U.et al. (1991). Neurogenic hyperalgesia: central neural correlates in responses of spinothalamic tract neurons. J. Neurophysiol., 66, 228–246.CrossRefGoogle ScholarPubMed
Sjövall, H., Jodal, M. and Lundgren, O. (1987). Sympathetic control of intestinal fluid and electrolyte transport. News Physiol. Sci., 2, 214–217.Google Scholar
Skok, V. I. (1986). Spontaneous and reflex activities: general characteristics. In Autonomic and Enteric Ganglia, ed. Karczmar, K., Koketsu, K. and Nishi, S.. New York, London: Plenum Press, pp. 425–438.CrossRefGoogle Scholar
Skok, V. I. (2002). Nicotinic acetylcholine receptors in autonomic ganglia. Auton. Neurosci., 97, 1–11.CrossRefGoogle ScholarPubMed
Skok, V. I. and Ivanov, A. Y. (1983). What is the ongoing activity of sympathetic neurons?J. Auton. Nerv. Syst., 7, 263–270.CrossRefGoogle ScholarPubMed
Skok, V. I. and Ivanov, A. Y. (1987). Organization of presynaptic input to neurones of a sympathetic ganglion. In Organization of the Autonomic Nervous System: Central and Peripheral Mechanisms, ed. Ciriello, J., Calaresu, F. R., Renaud, L. P. and Polosa, C.. New York: Alan Liss, pp. 37–46.Google Scholar
Skok, V. I. and Ivanov, A. Y. (1989). ectectbehhaя aktиbhoctь beΓetatиbhьix ΓahΓлиeb [Natural activity in autonomic ganglia]. Kiev: Publisher Naukova Duma.
Sly, D. J., Colvill, L., McKinley, M. J. and Oldfield, B. J. (1999). Identification of neural projections from the forebrain to the kidney, using the virus pseudorabies. J. Auton. Nerv. Syst., 77, 73–82.CrossRefGoogle ScholarPubMed
Smeyne, R. J., Klein, R., Schnapp, A.et al. (1994). Severe sensory and sympathetic neuropathies in mice carrying a disrupted Tr/nerve growth factor receptor gene. Nature, 368, 246–249.CrossRefGoogle Scholar
Smith, J. C., Ellenberger, H. H., Ballanyi, K., Richter, D. W. and Feldman, J. L. (1991). Pre-Bötzinger complex: a brain stem region that may generate respiratory rhythm in mammals. Science, 254, 726–729.CrossRefGoogle ScholarPubMed
Smith, J. C., Butera, R. J., Koshiya, N.et al. (2000). Respiratory rhythm generation in neonatal and adult mammals: the hybrid pacemaker-network model. Respir. Physiol., 122, 131–147.CrossRefGoogle ScholarPubMed
Smith, J. E., Jansen, A. S., Gilbey, M. P. and Loewy, A. D. (1998). central nervous system cell groups projecting to sympathetic outflow of tail artery: neural circuits involved in heat loss in the rat. Brain Res., 786, 153–164.CrossRefGoogle ScholarPubMed
Smith, O. A. and DeVito, J. L. (1984). Central neural integration for the control of autonomic responses associated with emotion. Annu. Rev. Neurosci., 7, 43–65.CrossRefGoogle ScholarPubMed
Smith, O. A., Hohimer, A. R., Astley, C. A. and Taylor, D. J. (1979). Renal and hindlimb vascular control during acute emotion in the baboon. Am. J. Physiol., 236, R198–R205.Google ScholarPubMed
Smith, O. A., Astley, C. A., DeVito, J. L., Stein, J. M. and Walsh, K. E. (1980). Functional analysis of hypothalamic control of the cardiovascular responses accompanying emotional behavior. Fed. Proc., 39, 2487–2494.Google ScholarPubMed
Smith, O. A., Astley, C. A., Spelman, F. A.et al. (2000). Cardiovascular responses in anticipation of changes in posture and locomotion. Brain Res. Bull., 53, 69–76.CrossRefGoogle ScholarPubMed
Smith, R. E. and Horwitz, B. A. (1969). Brown fat and thermogenesis. Physiol. Rev., 49, 330–425.CrossRefGoogle ScholarPubMed
Smith, T. K., Oliver, G. R., Hennig, G. W.et al. (2003). A smooth muscle tone-dependent stretch-activated migrating motor pattern in isolated guinea-pig distal colon. J. Physiol., 551, 955–969.CrossRefGoogle ScholarPubMed
Söderholm, J. D. and Perdue, M. H. (2001). Stress and gastrointestinal tract. II. Stress and intestinal barrier function. Am. J. Physiol. Gastrointest. Liver Physiol., 280, G7–G13.CrossRefGoogle ScholarPubMed
Sonnenschein, R. R. and Weissman, M. L. (1978). Sympathetic vasomotor outflows to hindlimb muscles of the cat. Am. J. Physiol., 235, H482–H487.Google ScholarPubMed
Spencer, N. J. and Smith, T. K. (2001). Simultaneous intracellular recordings from longitudinal and circular muscle during the peristaltic reflex in guinea-pig distal colon. J. Physiol., 533, 787–799.CrossRefGoogle ScholarPubMed
Spencer, N. J. and Smith, T. K. (2004). Mechanosensory S-neurons rather than AH-neurons appear to generate a rhythmic motor pattern in guinea-pig distal colon. J. Physiol., 558, 577–596.CrossRefGoogle ScholarPubMed
Spencer, N., Walsh, M. and Smith, T. K. (1999). Does the guinea-pig ileum obey the ‘law of the intestine’? J. Physiol., 517, 889–898.CrossRefGoogle ScholarPubMed
Spencer, N. J., Hennig, G. W. and Smith, T. K. (2002). A rhythmic motor pattern activated by circumferential stretch in guinea-pig distal colon. J. Physiol., 545, 629–648.CrossRefGoogle ScholarPubMed
Spencer, N. J., Hennig, G. W. and Smith, T. K. (2003a). Stretch-activated neuronal pathways to longitudinal and circular muscle in guinea pig distal colon. Am. J. Physiol. Gastrointest. Liver Physiol., 284, G231–G241.CrossRefGoogle Scholar
Spencer, N. J., Sanders, K. M. and Smith, T. K. (2003b). Migrating motor complexes do not require electrical slow waves in the mouse small intestine. J. Physiol., 553, 881–893.CrossRefGoogle Scholar
Spencer, S. E., Sawyer, W. B., Wada, H., Platt, K. B. and Loewy, A. D. (1990). central nervous system projections to the pterygopalatine parasympathetic preganglionic neurons in the rat: a retrograde transneuronal viral cell body labeling study. Brain Res., 534, 149–169.CrossRefGoogle ScholarPubMed
Spike, R. C., Puskar, Z., Andrew, D. and Todd, A. J. (2003). A quantitative and morphological study of projection neurons in lamina I of the rat lumbar spinal cord. Eur. J. Neurosci., 18, 2433–2448.CrossRefGoogle ScholarPubMed
Spillane, D. M. (1981). The Doctrine of the Nerves. Oxford: Oxford University Press.Google Scholar
Spyer, K. M. (1981). Neural organisation and control of the baroreceptor reflex. Rev. Physiol. Biochem. Pharmacol., 88, 24–124.Google ScholarPubMed
Spyer, K. M. (1994). Central nervous mechanisms contributing to cardiovascular control. J. Physiol., 474, 1–19.CrossRefGoogle ScholarPubMed
Squire, L. R., Bloom, F. E., McConnell, S. R.et al. (eds.) (2003). Fundamental Neuroscience, 2nd edn. San Diego: Academic Press.Google Scholar
St.-John, W. M. (1998). Neurogenesis of patterns of automatic ventilatory activity. Prog. Neurobiol., 56, 97–117.CrossRefGoogle ScholarPubMed
St.-John, W. M. and Paton, J. F. (2003). Defining eupnea. Respir. Physiol. Neurobiol., 139, 97–103.CrossRefGoogle ScholarPubMed
St.-John, W. M. and Paton, J. F. (2004). Role of pontile mechanisms in the neurogenesis of eupnea. Respir. Physiol. Neurobiol., 143, 321–332.CrossRefGoogle ScholarPubMed
Stanton-Hicks, M., Jänig, W., Hassenbusch, S.et al. (1995). Reflex sympathetic dystrophy: changing concepts and taxonomy. Pain, 63, 127–133.CrossRefGoogle ScholarPubMed
Staras, K., Chang, H. S. and Gilbey, M. P. (2001). Resetting of sympathetic rhythm by somatic afferents causes post-reflex coordination of sympathetic activity in rat. J. Physiol., 533, 537–545.CrossRefGoogle ScholarPubMed
Stein, R. D. and Weaver, L. C. (1988). Multi- and single-fibre mesenteric and renal sympathetic responses to chemical stimulation of intestinal receptors in cats. J. Physiol., 396, 155–172.CrossRefGoogle ScholarPubMed
Sterling, P. and Eyer, J. (1988). Allostasis: a new paradigm to explain arousal pathology. In Handbook of Life Stress, Cognition and Health, ed. Fisher, S. and Reason, J.. New York: Wiley, pp. 629–649.Google Scholar
Stjernberg, L. and Wallin, B. G. (1983). Sympathetic neural outflow in spinal man. A preliminary report. J. Auton. Nerv. Syst., 7, 313–318.CrossRefGoogle Scholar
Stjernberg, L., Blumberg, H. and Wallin, B. G. (1986). Sympathetic activity in man after spinal cord injury. Outflow to muscle below the lesion. Brain, 109, 695–715.CrossRefGoogle ScholarPubMed
Stornetta, R. L., Schreihofer, A. M., Pelaez, N. M., Sevigny, C. P. and Guyenet, P. G. (2001). Preproenkephalin messenger ribonucleic acid is expressed by C1 and non-C1 barosensitive bulbospinal neurons in the rostral ventrolateral medulla of the rat. J. Comp. Neurol., 435, 111–126.CrossRefGoogle ScholarPubMed
Stornetta, R. L., Sevigny, C. P., Schreihofer, A. M., Rosin, D. L. and Guyenet, P. G. (2002). Vesicular glutamate transporter DNPI/VGLUT2 is expressed by both C1 adrenergic and nonaminergic presympathetic vasomotor neurons of the rat medulla. J. Comp. Neurol., 444, 207–220.CrossRefGoogle ScholarPubMed
Stornetta, R. L., McQuiston, T. J. and Guyenet, P. G. (2004). γ-aminobutyric acidergic and glycinergic presympathetic neurons of rat medulla oblongata identified by retrograde transport of pseudorabies virus and in situ hybridization. J. Comp. Neurol., 479, 257–270.CrossRefGoogle Scholar
Strack, A. M. and Loewy, A. D. (1990). Pseudorabies virus: a highly specific transneuronal cell body marker in the sympathetic nervous system. J. Neurosci., 10, 2139–2147.CrossRefGoogle ScholarPubMed
Strack, A. M., Sawyer, W. B., Marubio, L. M. and Loewy, A. D. (1988). Spinal origin of sympathetic preganglionic neurons in the rat. Brain Res., 455, 187–191.CrossRefGoogle ScholarPubMed
Strack, A. M., Sawyer, W. B., Hughes, J. H., Platt, K. B. and Loewy, A. D. (1989a). A general pattern of central nervous system innervation of the sympathetic outflow demonstrated by transneuronal pseudorabies viral infections. Brain Res., 491, 156–162.CrossRefGoogle Scholar
Strack, A. M., Sawyer, W. B., Platt, K. B. and Loewy, A. D. (1989b). central nervous system cell groups regulating the sympathetic outflow to adrenal gland as revealed by transneuronal cell body labeling with pseudorabies virus. Brain Res., 491, 274–296.CrossRefGoogle Scholar
Strauther, G. R., Longo, W. E., Virgo, K. S. and Johnson, F. E. (1999). Appendicitis in patients with previous spinal cord injury. Am. J. Surg., 178, 403–405.CrossRefGoogle ScholarPubMed
Stricker, E. M. and Verbalis, J. G. (2003). Water intake and body fluids. In Fundamental Neuroscience, 2nd edn., ed. Squire, L. R., Bloom, F. E., McConnell, S. K.et al. San Diego: Academic Press, pp. 1011–1029.Google Scholar
Strickland, J. H. and Calhoun, M. L. (1963). The integumentary system of the cat. Am. J. Vet. Res., 24, 1019–1029.Google ScholarPubMed
Su, X. and Gebhart, G. F. (1998). Mechanosensitive pelvic nerve afferent fibers innervating the colon of the rat are polymodal in character. J. Neurophysiol., 80, 2632–2644.CrossRefGoogle ScholarPubMed
Su, X., Sengupta, J. N. and Gebhart, G. F. (1997a). Effects of kappa opioid receptor-selective agonists on responses of pelvic nerve afferents to noxious colorectal distension. J. Neurophysiol., 78, 1003–1012.CrossRefGoogle Scholar
Su, X., Sengupta, J. N. and Gebhart, G. F. (1997b). Effects of opioids on mechanosensitive pelvic nerve afferent fibers innervating the urinary bladder of the bat. J. Neurophysiol., 77, 1566–1580.CrossRefGoogle Scholar
Sugenoya, J., Iwase, S., Mano, T.et al. (1998). Vasodilator component in sympathetic nerve activity destined for the skin of the dorsal foot of mildly heated humans. J. Physiol., 507, 603–610.CrossRefGoogle ScholarPubMed
Sugiura, Y., Terui, N. and Hosoa, Y. (1989). Differences in the distribution of central terminals between visceral and somatic unmyelinated primary afferent fibers. J. Neurophysiol., 62, 834–847.CrossRefGoogle Scholar
Sun, M. K. (1995). Central neural organization and control of sympathetic nervous system in mammals. Prog. Neurobiol., 47, 157–233.CrossRefGoogle ScholarPubMed
Sun, M. K. and Guyenet, P. G. (1985). γ-aminobutyric acid-mediated baroreceptor inhibition of reticulospinal neurons. Am. J. Physiol., 249, R672–R680.Google Scholar
Sun, M. K. and Reis, D. J. (1996). Medullary vasomotor activity and hypoxic sympathoexcitation in pentobarbital-anesthetized rats. Am. J. Physiol., 270, R348–R355.Google ScholarPubMed
Sun, M. K. and Spyer, K. M. (1991). Nociceptive inputs into rostral ventrolateral medulla-spinal vasomotor neurones in rats. J. Physiol., 436, 685–700.CrossRefGoogle ScholarPubMed
Sun, M. K., Hackett, J. T. and Guyenet, P. G. (1988a). Sympathoexcitatory neurons of rostral ventrolateral medulla exhibit pacemaker properties in the presence of a glutamate-receptor antagonist. Brain Res., 438, 23–40.CrossRefGoogle Scholar
Sun, M. K., Young, B. S., Hackett, J. T. and Guyenet, P. G. (1988b). Reticulospinal pacemaker neurons of the rat rostral ventrolateral medulla with putative sympathoexcitatory function: an intracellular study in vitro. Brain Res., 442, 229–239.CrossRefGoogle Scholar
Sun, W. and Panneton, W. M. (2002). The caudal pressor area of the rat: its precise location and projections to the ventrolateral medulla. Am. J. Physiol. Regul. Integr. Comp. Physiol., 283, R768–R778.CrossRefGoogle ScholarPubMed
Sun, W. and Panneton, W. M. (2004). Defining projections from the caudal pressor area of the caudal ventrolateral medulla. J. Comp. Neurol., 482, 273–293.CrossRefGoogle Scholar
Sundlöf, G. and Wallin, B. G. (1977). The variability of muscle nerve sympathetic activity in resting recumbent man. J. Physiol., 272, 383–397.CrossRefGoogle ScholarPubMed
Sundlöf, G. and Wallin, B. G. (1978). Effect of lower body negative pressure on human muscle nerve sympathetic activity. J. Physiol., 278, 525–532.CrossRefGoogle ScholarPubMed
Sved, A. F. (2004). Tonic glutamatergic drive of rostral ventrolateral medulla vasomotor neurons? Am. J. Physiol. Regul. Integr. Comp. Physiol., 287, R1301–R1303.CrossRefGoogle Scholar
Sved, A. F., Mancini, D. L., Graham, J. C., Schreihofer, A. M. and Hoffman, G. E. (1994). PNMT-containing neurons of the C1 cell group express c-fos in response to changes in baroreceptor input. Am. J. Physiol., 266, R361–R367.Google ScholarPubMed
Swanson, L. W. (1987). The hypothalamus. In Handbook of Chemical Anatomy, Vol. 5, Integrated Systems of the central nervous system, part I: Hypothalamus, Hippocampus, Amygdala, Retina., ed. Björklund, A., Hökfelt, T. and Swanson, L. W., Amsterdam, New York, Oxford: Elsevier, pp. 1–124.Google Scholar
Swanson, L. W. (1995). Mapping the human brain: past, present, and future. Trends Neurosci., 18, 471–474.CrossRefGoogle ScholarPubMed
Swanson, L. W. (2000). Cerebral hemisphere regulation of motivated behavior. Brain Res., 886, 113–164.CrossRefGoogle ScholarPubMed
Swanson, L. W. (2003). The architecture of the nervous system. In Fundamental Neuroscience, 2nd edn., ed. Squire, L. R., Bloom, F. E., McConnell, S. K.et al. San Diego: Academic Press, pp. 15–45.Google Scholar
Swift, J. Q., Garry, M. G., Roszkowski, M. T. and Hargreaves, K. M. (1993). Effect of flurbiprofen on tissue levels on immunoreactive bradykinin and acute postoperative pain. J. Oral. Maxillofac. Surg., 51, 112–117.CrossRefGoogle ScholarPubMed
Szurszewski, J. H. (1981). Physiology of mammalian prevertebral ganglia. Annu. Rev. Physiol., 43, 53–68.CrossRefGoogle ScholarPubMed
Szurszewski, J. H. and King, B. F. (1989). Physiology of prevertebral ganglia in mammals with special reference to the inferior mesenteric ganglion. In The Gastrointestinal System, ed. Wood, J. D.. Bethesda: American Physiological Society, pp. 519–592.Google Scholar
Taché, Y., Martinez, V., Million, M. and Wang, L. (2001). Stress and the gastrointestinal tract III. Stress-related alterations of gut motor function: role of brain corticotropin-releasing factor receptors. Am. J. Physiol. Gastrointest. Liver Physiol., 280, G173–G177.CrossRefGoogle ScholarPubMed
Taiwo, Y. O. and Levine, J. D. (1988). Characterization of the arachidonic acid metabolites mediating bradykinin and noradrenaline hyperalgesia. Brain Res., 458, 402–406.CrossRefGoogle ScholarPubMed
Tanaka, M., Nagashima, K., McAllen, R. M. and Kanosue, K. (2002). Role of the medullary raphe in thermoregulatory vasomotor control in rats. J. Physiol., 540, 657–664.CrossRefGoogle ScholarPubMed
Tang, X., Neckel, N. D. and Schramm, L. P. (2003). Locations and morphologies of sympathetically correlated neurons in the T(10) spinal segment of the rat. Brain Res., 976, 185–193.CrossRefGoogle Scholar
Tang, X., Neckel, N. D. and Schramm, L. P. (2004). Spinal interneurons infected by renal injection of pseudorabies virus in the rat. Brain Res., 1004, 1–7.CrossRefGoogle ScholarPubMed
Tatarchenko, L. A., Ivanov, A. Y. and Skok, V. I. (1990). Organization of the tonically active pathways through the superior cervical ganglion of the rabbit. J. Auton. Nerv. Syst., 30 Suppl., S163–S168.CrossRefGoogle ScholarPubMed
Taylor, R. B. and Weaver, L. C. (1993). Dorsal root afferent influences on tonic firing of renal and mesenteric sympathetic nerves in rats. Am. J. Physiol., 264, R1193–R1199.Google ScholarPubMed
Horst, G. J., Hautvast, R. W., Jongste, M. J. and Korf, J. (1996). Neuroanatomy of cardiac activity-regulating circuitry: a transneuronal retrograde viral labelling study in the rat. Eur. J. Neurosci., 8, 2029–2041.CrossRefGoogle ScholarPubMed
Thompson, R. H. and Swanson, L. W. (2003). Structural characterization of a hypothalamic visceromotor pattern generator network. Brain Res. Brain Res. Rev., 41, 153–202.CrossRefGoogle ScholarPubMed
Thorén, P. (1979). Role of cardiac vagal C-fibers in cardiovascular control. Rev. Physiol. Biochem. Pharmacol., 86, 1–94.CrossRefGoogle ScholarPubMed
Thrasher, T. N. (2002). Unloading of arterial baroreceptors causes neurogenic hypertension. Am. J. Physiol. Regul. Integr. Comp. Physiol., 282, R1044–R1053.CrossRefGoogle ScholarPubMed
Thrasher, T. N. (2004). Baroreceptors and the long-term control of blood pressure. Exp. Physiol., 89, 331–335.CrossRefGoogle ScholarPubMed
Thrasher, T. N. (2005). Baroreceptors, baroreceptor unloading, and the long-term control of blood pressure. Am. J. Physiol. Integr. Comp. Physiol., 288, R819–R827.CrossRefGoogle ScholarPubMed
Thuneberg, L. (1982). Interstitial cells of Cajal: intestinal pacemaker cells? Adv. Anat. Embryol. Cell Biol., 71, 1–130.CrossRefGoogle ScholarPubMed
Thuneberg, L. (1999). One hundred years of interstitial cells of Cajal. Microsc. Res. Tech., 47, 223–238.3.0.CO;2-C>CrossRefGoogle ScholarPubMed
Tokimasa, T. and Akasu, T. (1995). Biochemical gating for voltage-gated channels: mechanisms for slow synaptic potentials in autonomic ganglia. In The Autonomic Nervous System, Vol. 6, Autonomic Ganglia, ed. McLachlan, E. M.. Chur Switzerland: Harwood Academic Publishers, pp. 259–295.Google Scholar
Tomita, T. (1975). Electrophysiology of mammalian smooth muscle. Prog. Biophys. Mol. Biol., 30, 185–203.CrossRefGoogle ScholarPubMed
Tomori, Z. and Widdicombe, J. G. (1969). Muscular, bronchomotor and cardiovascular reflexes elicited by mechanical stimulation of the respiratory tract. J. Physiol., 200, 25–49.CrossRefGoogle ScholarPubMed
Torebjörk, H. E., Lundberg, L. E. R. and LaMotte, R. H. (1992). Central changes in processing of mechanoreceptive input in capsaicin-induced secondary hyperalgesia in humans. J. Physiol., 448, 765–780.CrossRefGoogle ScholarPubMed
Traube, L. (1865). Über periodische Thätigkeits-Aeusserungen des vasomotorischen und Hemmungs-Nervencentrums [On the periodic changes of the vasomotor and inhibitory neural centers]. Centralbl. Medic. Wissensch., 56, 881–885.Google Scholar
Travagli, R. A. and Rogers, R. C. (2001). Receptors and transmission in the brain-gut axis: potential for novel therapies. V. Fast and slow extrinsic modulation of dorsal vagal complex circuits. Am. J. Physiol., 281, G595–G601.Google ScholarPubMed
Travagli, R. A., Hermann, G. E., Browning, K. N. and Rogers, C. (2006). Brain stem circuits regulating gastric function. Annu. Rev. Physiol., 68, 279–305.CrossRefGoogle ScholarPubMed
Treede, R. D., Kenshalo, D. R., Gracely, R. H. and Jones, A. K. (1999). The cortical representation of pain. Pain, 79, 105–111.CrossRefGoogle Scholar
Treede, R. D., Apkarian, A. V., Bromm, B., Greenspan, J. D. and Lenz, F. A. (2000). Cortical representation of pain: functional characterization of nociceptive areas near the lateral sulcus. Pain, 87, 113–119.CrossRefGoogle ScholarPubMed
Tsunoo, A., Konishi, S. and Otsuka, M. (1982). Substance P as an excitatory transmitter of primary afferent neurons in guinea-pig sympathetic ganglia. Neuroscience, 7, 2025–2037.CrossRefGoogle ScholarPubMed
Tucker, D. C., Saper, C. B., Ruggiero, D. A. and Reis, D. J. (1987). Organization of central adrenergic pathways: I. Relationships of ventrolateral medullary projections to the hypothalamus and spinal cord. J. Comp. Neurol., 259, 591–603.CrossRefGoogle ScholarPubMed
Ulman, L. G., Potter, E. K. and McCloskey, D. I. (1992). Effects of sympathetic activity and galanin on cardiac vagal action in anaesthetized cats. J. Physiol., 448, 225–235.CrossRefGoogle ScholarPubMed
Undem, B. and Weinreich, D. (eds.) (2005). Advance in Vagal Afferent Neurobiology. Boca Raton: CRC Press.CrossRefGoogle Scholar
Unzicker, K. (ed.) (1996). The Autonomic Nervous System Vol. 10, Autonomic–Endocrine Interactions. Amsterdam: Harwood Academic Publishers.Google Scholar
Uvnäs, B. (1954). Sympathetic vasodilator outflow. Pharmacol. Rev., 34, 608–618.Google ScholarPubMed
Uvnäs, B. (1960). Sympathetic vasodilator system and blood flow. Physiol. Rev., 40, Suppl. 4, 68–75.Google Scholar
Vallbo, A. B., Hagbarth, K. E., Torebjörk, H. E. and Wallin, B. G. (1979). Somatosensory, proprioceptive and sympathetic activity in human peripheral nerves. Physiol. Rev., 59, 919–957.CrossRefGoogle ScholarPubMed
Vallbo, A. B., Olsson, K. A., Westberg, K. G. and Clark, F. J. (1984). Microstimulation of single tactile afferents from the human hand. Sensory attributes related to unit type and properties of receptive fields. Brain, 107, 727–749.CrossRefGoogle ScholarPubMed
Vallbo, A. B., Hagbarth, K. E. and Wallin, B. G. (2004). Microneurography: how the technique developed and its role in the investigation of the sympathetic nervous system. J. Appl. Physiol., 96, 1262–1269.CrossRefGoogle ScholarPubMed
Bockstaele, E. J., Aston-Jones, G., Pieribone, V. A., Ennis, M. and Shipley, M. T. (1991). Subregions of the periaqueductal gray topographically innervate the rostral ventral medulla in the rat. J. Comp. Neurol., 309, 305–327.CrossRefGoogle ScholarPubMed
Dijk, J. G. (2003). Fainting in animals. Clin. Auton. Res., 13, 247–255.CrossRefGoogle ScholarPubMed
Helden, D. F. (1988a). An alpha-adrenoceptor-mediated chloride conductance in mesenteric veins of the guinea-pig. J. Physiol., 401, 489–501.CrossRefGoogle Scholar
Helden, D. F. (1988b). Electrophysiology of neuromuscular transmission in guinea-pig mesenteric veins. J. Physiol., 401, 469–488.CrossRefGoogle Scholar
Helden, D. F. (1991). Spontaneous and noradrenaline-induced transient depolarizations in the smooth muscle of guinea-pig mesenteric vein. J. Physiol., 437, 511–541.CrossRefGoogle ScholarPubMed
Vecchiet, L., Albe-Fessard, D., Lindblom, U. and Giamberardino, M. A. (eds.) (1993). New Trends in Referred Pain and Hyperalgesia. Amsterdam: Elsevier Science.Google Scholar
Verberne, A. J., Stornetta, R. L. and Guyenet, P. G. (1999). Properties of C1 and other ventrolateral medullary neurons with hypothalamic projections in the rat. J. Physiol., 517, 477–494.CrossRefGoogle ScholarPubMed
Victor, R. G., Leimbach, W. N. J., Seals, D. R., Wallin, B. G. and Mark, A. L. (1987). Effects of the cold pressor test on muscle sympathetic nerve activity in humans. Hypertension, 9, 429–436.CrossRefGoogle ScholarPubMed
Victor, R. G., Secher, N. H., Lyson, T. and Mitchell, J. H. (1995). Central command increases muscle sympathetic nerve activity during intense intermittent isometric exercise in humans. Circ. Res., 76, 127–131.CrossRefGoogle ScholarPubMed
Villanueva, L. and Nathan, P. W. (2000). Multiple pain pathways. In Proceedings of the 9th World Congress on Pain, ed. Devor, M., Rowbotham, M. C. and Weisenfeld-Hallin, Z. Seattle: IASP Press, pp. 371–386.Google Scholar
Vissing, S. F., Scherrer, U. and Victor, R. G. (1994). Increase of sympathetic discharge to skeletal muscle but not to skin during mild lower body negative pressure in humans. J. Physiol., 481, 233–241.CrossRefGoogle Scholar
Vizzard, M. A., Erickson, V. L., Card, J. P., Roppolo, J. R. and Groat, W. C. (1995). Transneuronal labeling of neurons in the adult rat brainstem and spinal cord after injection of pseudorabies virus into the urethra. J. Comp. Neurol., 355, 629–640.CrossRefGoogle ScholarPubMed
Vizzard, M. A., Brisson, M. and Groat, W. C. (2000). Transneuronal labeling of neurons in the adult rat central nervous system following inoculation of pseudorabies virus into the colon. Cell Tissue Res., 299, 9–26.CrossRefGoogle ScholarPubMed
Vogt, B. A. (2005). Pain and emotion interactions in subregions of the cingulate gyrus. Nat. Rev. Neurosci., 6, 533–544.CrossRefGoogle ScholarPubMed
Vogt, B. A., Rosene, D. L. and Pandya, D. N. (1979). Thalamic and cortical afferents differentiate anterior from posterior cingulate cortex in the monkey. Science, 204, 205–207.CrossRefGoogle ScholarPubMed
Vollmer, R. R. (1996). Selective neural regulation of epinephrine and norepinephrine cells in the adrenal medulla – cardiovascular implications. Clin. Exp. Hypertens., 18, 731–751.CrossRefGoogle ScholarPubMed
Euler, U. S. (1956). Noradrenaline: Chemistry, Physiology, Pharmacology and Clinical Aspects. Springfield Ill: Thomas.Google Scholar
von Euler, C. (1986). Brain stem mechanisms for generation and control of breathing pattern. In Handbook of Physiology, The Respiratory System, Vol. II, ed. Fishman, A. P., Cherniack, N. S. and Widdicombe, J. G.. Bethesda: American Physiological Society, pp. 1–67.Google Scholar
Holst, E. and St. Paul, U. (1960). Vom Wirkungsgefüge der Triebe [The Wirkungsgefüge of Drives]. Die Naturwissenschaften, 47, 409–422.CrossRefGoogle Scholar
Holst, E. and St. Paul, U. (1962). Electrically controlled behavior. Sci. Am., 206, 50–60.CrossRefGoogle Scholar
Voyvodic, J. T. (1989). Peripheral target regulation of dendritic geometry in the rat superior cervical ganglion. J. Neurosci., 9, 1997–2010.CrossRefGoogle ScholarPubMed
Wallin, B. G. (1990). Neural control of human skin blood flow. J. Auton. Nerv. Syst., 30 Suppl., S185–S190.CrossRefGoogle ScholarPubMed
Wallin, B. G. (2002). Intraneural recordings of normal and abnormal sympathetic activity in humans. In Autonomic Failure, 4th edn., ed. Mathias, C. J. and Bannister, R.. Oxford: Oxford University Press, pp. 224–231.Google ScholarPubMed
Wallin, B. G. and Elam, M. (1994). Insights from intraneural recordings of sympathetic nerve traffic in humans. News Physiol. Sci., 9, 203–207.Google Scholar
Wallin, B. G. and Fagius, J. (1986). The sympathetic nervous system in man – aspects derived from microelectrode recordings. Trends Neurosci., 9, 63–67.CrossRefGoogle Scholar
Wallin, B. G. and Fagius, J. (1988). Peripheral sympathetic neural activity in conscious humans. Annu. Rev. Physiol., 50, 565–576.CrossRefGoogle ScholarPubMed
Wallin, B. G. and Stjernberg, L. (1984). Sympathetic activity in man after spinal cord injury. Outflow to skin below the lesion. Brain, 107, 183–198.CrossRefGoogle ScholarPubMed
Wallin, B. G., Batelsson, K., Kienbaum, P., Karlsson, T. and Gazelius, B. (1998). Two neural mechanisms for respiration-induced cutaneous vasodilatation in humans. J. Physiol., 513, 559–569.CrossRefGoogle ScholarPubMed
Wang, F. B., Holst, M. C. and Powley, T. L. (1995). The ratio of pre- to postganglionic neurons and related issues in the autonomic nervous system. Brain Res. Brain Res. Rev., 21, 93–115.CrossRefGoogle ScholarPubMed
Wang, H., Germanson, T. P. and Guyenet, P. G. (2002). Depressor and tachypneic responses to chemical stimulation of the ventral respiratory group are reduced by ablation of neurokinin-1 receptor-expressing neurons. J. Neurosci., 22, 3755–3764.CrossRefGoogle ScholarPubMed
Wang, H., Weston, M. C., McQuiston, T. J., Stornetta, R. L. and Guyenet, P. G. (2003). Neurokinin-1 receptor-expressing cells regulate depressor region of rat ventrolateral medulla. Am. J. Physiol. Heart Circ. Physiol., 285, H2757–H2769.CrossRefGoogle ScholarPubMed
Wank, M. and Neuhuber, W. L. (2001). Local differences in vagal afferent innervation of the rat esophagus are reflected by neurochemical differences at the level of the sensory ganglia and by different brainstem projections. J. Comp. Neurol., 435, 41–59.CrossRefGoogle ScholarPubMed
Ward, S. M. and Sanders, K. M. (2001). Physiology and pathophysiology of the interstitial cell of Cajal: from bench to bedside. I. Functional development and plasticity of interstitial cells of Cajal networks. Am. J. Physiol. Gastrointest. Liver Physiol., 281, G602–G611.CrossRefGoogle ScholarPubMed
Ward, S. M., Morris, G., Reese, L., Wang, X. Y. and Sanders, K. M. (1998). Interstitial cells of Cajal mediate enteric inhibitory neurotransmission in the lower esophageal and pyloric sphincters. Gastroenterology, 115, 314–329.CrossRefGoogle ScholarPubMed
Ward, S. M., Beckett, E. A., Wang, X.et al. (2000a). Interstitial cells of Cajal mediate cholinergic neurotransmission from enteric motor neurons. J. Neurosci., 20, 1393–1403.CrossRefGoogle Scholar
Ward, S. M., Ordog, T., Koh, S. D.et al. (2000b). Pacemaking in interstitial cells of Cajal depends upon calcium handling by endoplasmic reticulum and mitochondria. J. Physiol., 525, 355–361.CrossRefGoogle Scholar
Wasner, G., Heckmann, K., Maier, C. and Baron, R. (1999). Vascular abnormalities in acute reflex sympathetic dystrophy (complex regional pain syndrome I): complete inhibition of sympathetic nerve activity with recovery. Arch. Neurol., 56, 613–620.CrossRefGoogle ScholarPubMed
Wasner, G., Schattschneider, J., Heckmann, K., Maier, C. and Baron, R. (2001). Vascular abnormalities in reflex sympathetic dystrophy (complex regional pain syndrome I): mechanisms and diagnostic value. Brain, 124, 587–599.CrossRefGoogle ScholarPubMed
Watkins, L. R. and Maier, S. F. (eds.) (1999). Cytokines and Pain. Basel, Boston, Berlin: Birkhäuser Verlag.CrossRefGoogle ScholarPubMed
Watkins, L. R. and Maier, S. F. (2000). The pain of being sick: implications of immune-to-brain communication for understanding pain. Annu. Rev. Psychol., 51, 29–57.CrossRefGoogle ScholarPubMed
Watts, A. G. and Swanson, L. W. (2002). Anatomy of motivational systems. In “Stevens” Handbook of Experimental Psychology, 3rd edn. Vol. 3, ed. Gallistel, G. R.. New York: John Wiley, pp. 563–631.Google Scholar
Weaver, L. C. and Polosa, C. (1997). Spinal cord circuits providing control of sympathetic preganglionic neurons. In The Autonomic Nervous System, Vol. 11, Central Nervous Control of Autonomic Function, ed. Jordan, D.. Amsterdam: Harwood Academic Publishers, pp. 29–61.Google Scholar
Weaver, L. C., Verghese, P., Bruce, J. C.et al. (2001). Autonomic dysreflexia and primary afferent sprouting after clip-compression injury of the rat spinal cord. J. Neurotrauma, 18, 1107–1119.CrossRefGoogle ScholarPubMed
Weaver, L. C., Marsh, D. R., Gris, D., Meakin, S. O. and Dekaban, G. A. (2002). Central mechanisms for autonomic dysreflexia after spinal cord injury. Prog. Brain Res., 137, 83–95.CrossRefGoogle ScholarPubMed
Wesselmann, U. and McLachlan, E. M. (1984). The effect of previous transection on quantitative estimates of the preganglionic neurons projecting in the cervical sympathetic trunk of the guinea-pig and the cat made by retrograde labelling of damaged axons by horseradish peroxidase. Neuroscience, 13, 1299–1309.CrossRefGoogle Scholar
Westerhaus, M. J. and Loewy, A. D. (1999). Sympathetic-related neurons in the preoptic region of the rat identified by viral transneuronal labeling. J. Comp. Neurol., 414, 361–378.3.0.CO;2-X>CrossRefGoogle ScholarPubMed
Westerhaus, M. J. and Loewy, A. D. (2001). Central representation of the sympathetic nervous system in the cerebral cortex. Brain Res., 903, 117–127.CrossRefGoogle ScholarPubMed
Weston, M., Wang, H., Stornetta, R. L., Sevigny, C. P. and Guyenet, P. G. (2003). Fos expression by glutamatergic neurons of the solitary tract nucleus after phenylephrine-induced hypertension in rats. J. Comp. Neurol., 460, 525–541.CrossRefGoogle ScholarPubMed
White, J. C. and Bland, E. F. (1948). The surgical relief of severe angina pectoris. Medicine, 27, 1–42.CrossRefGoogle ScholarPubMed
Whyment, A. D., Wilson, J. M., Renaud, L. P. and Spanswick, D. (2004). Activation and integration of bilateral γ-aminobutyric acid-mediated synaptic inputs in neonatal rat sympathetic preganglionic neurons in vitro. J. Physiol., 555, 189–203.CrossRefGoogle Scholar
Widdicombe, J. G. (1966). Action potentials in parasympathetic and sympathetic efferent fibers to the trachea and lungs of dogs and cats. J. Physiol., 186, 56–88.CrossRefGoogle Scholar
Widdicombe, J. G. (1986). Sensory innervation of the lungs and airways. Prog. Brain Res., 67, 49–64.CrossRefGoogle ScholarPubMed
Widdicombe, J. (2001). Airway receptors. Respir. Physiol., 125, 3–15.CrossRefGoogle ScholarPubMed
Widdicombe, J. (2003). Functional morphology and physiology of pulmonary rapidly adapting receptors (RARs). Anat. Rec., 270A, 2–10.CrossRefGoogle Scholar
Williams, C. L., Men, D., Clayton, E. C. and Gold, P. E. (1998). Norepinephrine release in the amygdala after systemic injection of epinephrine or escapable footshock: contribution of the nucleus of the solitary tract. Behav. Neurosci., 112, 1414–1422.CrossRefGoogle ScholarPubMed
Williams, I. R. and Kupper, T. S. (1996). Immunity at the surface: homeostatic mechanisms of the skin immune system. Life Sci., 58, 1485–1507.CrossRefGoogle ScholarPubMed
Williams, R. M., Berthoud, H. R. and Stead, R. H. (1997). Vagal afferent nerve fibers contact mast cells in rat small intestinal mucosa. NeuroImmunoModulation, 4, 266–270.CrossRefGoogle ScholarPubMed
Willis, W. D. J. (1985). The Pain System. Basel: Karger.Google ScholarPubMed
Willis, W. D. Jr. (1999). Dorsal root potentials and dorsal root reflexes: a double-edged sword. Exp. Brain Res., 124, 395–421.CrossRefGoogle ScholarPubMed
Willis, W. D. J. and Coggeshall, R. E. (2004a). Sensory Mechanisms of the Spinal Cord. Primary Afferent Neurons and the Spinal Dorsal Horn, Vol. 1, 3rd edn. New York: Kluwer Academic/Plenum Publishers.Google Scholar
Willis, W. D. J. and Coggeshall, R. E. (2004b). Sensory Mechanisms of the Spinal Cord. Ascending Sensory Tracts and Their Descending Control, Vol. 2, 3rd edn. New York: Kluwer Academic/Plenum Publishers.Google Scholar
Willis, W. D., Al Chaer, E. D., Quast, M. J. and Westlund, K. N. (1999). A visceral pain pathway in the dorsal column of the spinal cord. Proc. Natl. Acad. Sci. U. S. A., 96, 7675–7679.CrossRefGoogle ScholarPubMed
Willis, W. D. Jr., Zhang, X., Honda, C. N. and Giesler, G. J. Jr. (2001). Projections from the marginal zone and deep dorsal horn to the ventrobasal nuclei of the primate thalamus. Pain, 92, 267–276.CrossRefGoogle ScholarPubMed
Willis, W. D. Jr., Zhang, X., Honda, C. N. and Giesler, G. J. Jr. (2002). A critical review of the role of the proposed posterior part of the ventromedial nucleus (thalamus, primate) nucleus in pain. J. Pain, 3, 79–94.CrossRefGoogle Scholar
Wilson, L. B., Andrew, D. and Craig, A. D. (2002). Activation of spinobulbar lamina I neurons by static muscle contraction. J. Neurophysiol., 87, 1641–1645.CrossRefGoogle ScholarPubMed
Winkler, H. (1988). Occurrence and mechanisms of exocytosis in adrenal medulla and sympathetic nerve. In Handbook of Experimental Pharmacology, Vol. 90/I Catecholamines I, ed. Trendelenburg, U. and Weiner, N.. Berlin, Heidelberg, New York: Springer-Verlag, pp. 43–118.Google Scholar
Wood, J. D. (1994). Physiology of the enteric nervous system. In Physiology of the Gastrointestinal Tract, 3rd edn., ed. Johnson, L. R.. New York: Raven Press, pp. 423–482.Google Scholar
Wood, J. D. (2002). Enteric neuro-immunology. In The Autonomic Nervous System, Vol. 14, Innervation of the Gastrointestinal Tract, ed. Brookes, S. and Costa, M.. London, New York: Taylor and Francis, pp. 363–392.Google Scholar
Woods, S. C. and Stricker, E. M. (2003). Food intake and metabolism. In Fundamental Neuroscience, 2nd edn., eds. Squire, L. R., Bloom, F. E., McConnell, S. K.et al. San Diego: Academic Press, pp. 991–1009.Google Scholar
Woolf, C. J. (1996). Phenotypic modification of primary sensory neurons: the role of nerve growth factor in the production of persistent pain. Phil. Trans. R. Soc. London B, 351, 441–448.CrossRefGoogle ScholarPubMed
Woolf, C. J., Safieh-Garabedian, B., Ma, Q.-P., Crilly, P. and Winter, J. (1994). Nerve growth factor contributes to the generation of inflammatory sensory hypersensitivity. Neuroscience, 62, 327–331.CrossRefGoogle ScholarPubMed
Woolf, C. J., Ma, Q. P., Allchorne, A. and Poole, S. (1996). Peripheral cell types contributing to the hyperalgesic action of nerve growth factor in inflammation. J. Neurosci., 16, 2716–2723.CrossRefGoogle ScholarPubMed
Xi, X., Randall, W. C. and Wurster, R. D. (1991). Intracellular recording of spontaneous activity of canine intracardiac ganglion cells. Neurosci Lett., 128, 129–132.CrossRefGoogle ScholarPubMed
Yang, H., Kawakubo, K., Wong, H.et al. (2000a). Peripheral PYY inhibits intracisternal thyrotropin-releasing hormone-induced gastric acid secretion by acting in the brain. Am. J. Physiol. Gastrointest. Liver Physiol., 279, G575–G581.CrossRefGoogle Scholar
Yang, H., Yuan, P. Q., Wang, L. and Taché, Y. (2000b). Activation of the parapyramidal region in the ventral medulla stimulates gastric acid secretion through vagal pathways in rats. Neuroscience, 95, 773–779.CrossRefGoogle Scholar
Yang, Z. and Coote, J. H. (2003). Role of γ-aminobutyric acid and nitric oxide in the paraventricular nucleus-mediated reflex inhibition of renal sympathetic nerve activity following stimulation of right atrial receptors in the rat. Exp. Physiol., 88. 335–342.CrossRefGoogle ScholarPubMed
Yang, Z., Smith, L. and Coote, J. H. (2004). Paraventricular nucleus activation of renal sympathetic neurones is synaptically depressed by nitric oxide and glycine acting at a spinal level. Neuroscience, 124, 421–428.CrossRefGoogle Scholar
Yeoh, M., McLachlan, E. M. and Brock, J. A. (2004a). Tail arteries from chronically spinal rats have potentiated responses to nerve stimulation in vitro. J. Physiol., 556, 545–555.CrossRefGoogle Scholar
Yeoh, M., McLachlan, E. M. and Brock, J. A. (2004b). Chronic decentralization potentiates neurovascular transmission in the isolated rat tail artery, mimicking the effects of spinal transection. J. Physiol., 561, 583–596.CrossRefGoogle Scholar
Yoshimura, M., Polosa, C. and Nishi, S. (1987a). Slow excitatory postsynaptic potential and the depolarizing action of noradrenaline on sympathetic preganglionic neurons. Brain Res., 414, 138–142.CrossRefGoogle Scholar
Yoshimura, M., Polosa, C. and Nishi, S. (1987b). Slow inhibitory postsynaptic potential and the noradrenaline-induced inhibition of the cat sympathetic preganglionic neuron in vitro. Brain Res., 419, 383–386.CrossRefGoogle Scholar
Young, J. B. and Landsberg, L. (2001). Synthesis, storage, and secretion of adrenal medullary hormones: physiology and pathophysiology. In Handbook of Physiology. Section 7. The Endocrine System. Vol. IV, Coping with the Environment: Neural and Neuroendocrine Mechanisms, ed. McEwan, B. S.. Oxford, New York: Oxford University Press, pp. 3–19.Google Scholar
Yu, L. C. and Perdue, M. H. (2001). Role of mast cells in intestinal mucosal function: studies in models of hypersensitivity and stress. Immunol. Rev., 179, 61–73.CrossRefGoogle Scholar
Zagon, A. and Smith, A. D. (1993). Monosynaptic projections from the rostral ventrolateral medulla oblongata to identified sympathetic preganglionic neurons. Neuroscience, 54, 729–743.CrossRefGoogle ScholarPubMed
Zagorodnyuk, V. P. and Brookes, S. J. (2000). Transduction sites of vagal mechanoreceptors in the guinea pig esophagus. J. Neurosci., 20, 6249–6255.CrossRefGoogle ScholarPubMed
Zagorodnyuk, V. P., Chen, B. N. and Brookes, S. J. (2001). Intraganglionic laminar endings are mechano-transduction sites of vagal tension receptors in the guinea-pig stomach. J. Physiol., 534, 255–268.CrossRefGoogle ScholarPubMed
Zaretsky, D. V., Zaretskaia, M. V. and DiMicco, J. A. (2003a). Stimulation and blockade of γ-aminobutyric acid(A) receptors in the raphe pallidus: effects on body temperature, heart rate, and blood pressure in conscious rats. Am. J. Physiol. Regul. Integr. Comp. Physiol., 285, R110–R116.CrossRefGoogle Scholar
Zaretsky, D. V., Zaretskaia, M. V., Samuels, B. C., Cluxton, L. K. and DiMicco, J. A. (2003b). Microinjection of muscimol into raphe pallidus suppresses tachycardia associated with air stress in conscious rats. J. Physiol., 546, 243–250.CrossRefGoogle Scholar
Zhang, X. and Giesler, G. J. Jr. (2005). Response characterstics of spinothalamic tract neurons that project to the posterior thalamus in rats. J. Neurophysiol., 93, 2552–2564.CrossRefGoogle ScholarPubMed
Zhang, X., Fogel, R. and Renehan, W. E. (1992). Physiology and morphology of neurons in the dorsal motor nucleus of the vagus and the nucleus of the solitary tract that are sensitive to distension of the small intestine. J. Comp. Neurol., 323, 432–448.CrossRefGoogle ScholarPubMed
Zhang, X., Fogel, R. and Renehan, W. E. (1995a). Relationships between the morphology and function of gastric- and intestine-sensitive neurons in the nucleus of the solitary tract. J. Comp. Neurol., 363, 37–52.CrossRefGoogle Scholar
Zhang, X., Kostarczyk, E. and Giesler, G. J. Jr. (1995b). Spinohypothalamic tract neurons in the cervical enlargement of rats: descending axons in the ipsilateral brain. J. Neurosci., 15, 8393–8407.CrossRefGoogle Scholar
Zhang, X., Renehan, W. E. and Fogel, R. (1998). Neurons in the vagal complex of the rat respond to mechanical and chemical stimulation of the GI tract. Am. J. Physiol., 274, G331–G341.Google ScholarPubMed
Zhang, X., Honda, C. N. and Giesler, G. J. Jr. (2000a). Position of spinothalamic tract axons in upper cervical spinal cord of monkeys. J. Neurophysiol., 84, 1180–1185.CrossRefGoogle Scholar
Zhang, X., Wenk, H. N., Honda, C. N. and Giesler, G. J. Jr. (2000b). Locations of spinothalamic tract axons in cervical and thoracic spinal cord white matter in monkeys. J. Neurophysiol., 83, 2869–2880.CrossRefGoogle Scholar
Zhao, F. Y., Saito, K., Yoshioka, K.et al. (1996). Tachykininergic synaptic transmission in the coeliac ganglion of the guinea-pig. Br. J. Pharmacol., 118, 2059–2066.CrossRefGoogle ScholarPubMed
Zhu, J. X., Zhu, X. Y., Owyang, C. and Li, Y. (2001). Intestinal serotonin acts as a paracrine substance to mediate vagal signal transmission evoked by luminal factors in the rat. J. Physiol., 530, 431–442.CrossRefGoogle ScholarPubMed

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.

  • References
  • Wilfrid Jänig, Christian-Albrechts Universität zu Kiel, Germany
  • Book: Integrative Action of the Autonomic Nervous System
  • Online publication: 10 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511541667.017
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.

  • References
  • Wilfrid Jänig, Christian-Albrechts Universität zu Kiel, Germany
  • Book: Integrative Action of the Autonomic Nervous System
  • Online publication: 10 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511541667.017
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.

  • References
  • Wilfrid Jänig, Christian-Albrechts Universität zu Kiel, Germany
  • Book: Integrative Action of the Autonomic Nervous System
  • Online publication: 10 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511541667.017
Available formats
×