Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-19T14:42:42.216Z Has data issue: false hasContentIssue false

Bibliography

Published online by Cambridge University Press:  29 November 2018

Donald Pfaff
Affiliation:
Neurobiology and Behavior, Rockefeller University, New York
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
How Brain Arousal Mechanisms Work
Paths Toward Consciousness
, pp. 123 - 152
Publisher: Cambridge University Press
Print publication year: 2018

Access options

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

References

Adamantidis, A., Zhang, F., Aravanis, A. M., Deisseroth, K., and de Lecea, L. (2007). Neural substrates of awakening probed with optogenetic control of hypocretin neurons. Nature. 450: 420–5.Google Scholar
Adams, J., Graham, D., Jennett, B. (2000). The neuropathology of the vegetative state after an acute brain insult. Brain. 123: 1327–38.CrossRefGoogle ScholarPubMed
Adams, Z. M., Forgacs, P. B., Conte, M. M., Nauvel, T. J., Drover, J. D. (2016). Late and progressive alterations of sleep dynamics following central thalamic deep brain stimulation (CT-DBS) in chronic minimally conscious state. Clin Neurophysiol. 127(9): 3086–92.Google Scholar
Agmo, A., Villalpando, A. (1995). Central nervous stimulants facilitate sexual behavior in male rats with medial prefrontal cortex lesions. Brain Res. 696(1–2): 187–93.Google Scholar
Albers, H. E. (2015). Species, sex and individual differences in the vasotocin/vasopressin system: relationship to neurochemical signaling in the social behavior neural network. Front Neuroendocrinol. 36: 4971.CrossRefGoogle ScholarPubMed
Albert, R., Jeong, H., Barabasi, A. L. (2000). Error and attack tolerance of complex networks. Nature. 406(6794): 378–82.Google Scholar
Albert, F. W., Carlborg, O., Plyusnina, I., et al. (2009). Genetic architecture of tameness in a rat model of animal domestication. Genetics. 182(2): 541–54. doi: 10.1534/genetics.109.102186. Epub 2009 Apr 10.Google Scholar
Allaway, K. C., Machold, R. (2017). Developmental specification of forebrain cholinergic neurons. Dev Biol. 421(1): 17.Google Scholar
Alonso, L. M., et al. (2014). Dynamical criticality during induction of anesthesia in human ECoG recordings. Front Neural Circuits. 8: 20. https://doi.org/10.3389/fncir.2014.00020.Google Scholar
Alvarez-Dieppa, A. C., et al. (2016). Vagus nerve stimulation enhances extinction of conditioned fear in rats and modulates arc protein, CaMKII, and GluN2B-containing NMDA receptors in the basolateral amygdala. Neural Plast. 2016:4273280.Google Scholar
Amaral, D. (2000). In Principles of Neural Science (4th edition). New York, NY: McGraw-Hill, pp. 331–53.Google Scholar
Anaclet, C., Parmentier, R., et al. (2009). Orexin/hypocretin and histamine: distinct roles in the control of wakefulness demonstrated using knock-out mouse models. J Neurosci. 29(46): 14423–38.Google Scholar
Anderson, D. J. (2016). Circuit modules linking internal states and social behaviour in flies and mice. Nat Rev Neurosci. 17(11): 692704.Google Scholar
Andrews-Hanna, J. R., et al. (2007). Disruption of large-scale brain systems in advanced aging. Neuron. 56(5): 924–35.Google Scholar
Anholt, R. R. (2004). Genetic modules and networks for behavior: lessons from Drosophila. Bioessays. 26(12): 1299–306.Google Scholar
Antony, A. K., Kong, W., Lorenz, H. P. (2010). Upregulation of neurodevelopmental genes during scarless healing. Ann Plast Surg. 64(2): 247–50.Google Scholar
Anthony, T. E., et al. (2014). Control of stress-induced persistent anxiety by an extra-amygdala septohypothalamic circuit. Cell. 156: 522–36.Google Scholar
Archer, J. (1977). Testosterone and persistence in mice. Anim Behav. 25(2): 479–88.Google Scholar
Arrieta-Cruz, I., Pfaff, D. W., Shelley, D. N. (2007). Mouse model of diffuse brain damage following anoxia, evaluated by assay of generalized arousal. Exp Neurol. 205(2): 449–60.Google Scholar
Aston-Jones, G., Cohen, J. D. (2005). An integrative theory of locus coeruleus-norepinephrine function: adaptive gain and optimal performance. Ann Rev Neurosci. 28: 403–50.Google Scholar
Aston-Jones, G., Zhu, Y., Card, J. P. (2004). Numerous GABAergic afferents to locus ceruleus in the pericerulear dendritic zone: possible interneuronal pool. J Neurosci. 24(9): 2313–21.Google Scholar
Aston-Jones, G., et al. (1996). Role of the locus coeruleus in emotional activation. Prog Brain Res. 107: 379402.Google Scholar
Aston-Jones, G., et al. (2001). A neural circuit for circadian regulation of arousal. Nat Neurosci. 4: 732–8.Google Scholar
Autret, A., Lucas, B., Mondon, K., et al. (2001). Sleep and brain lesions: a critical review of the literature and additional new cases. Neurophysiol Clin. 31: 356–75.CrossRefGoogle ScholarPubMed
Azevedo, R. T., Badoud, D., Tsakiris, M. (2017). Afferent cardiac signals modulate attentional engagement to low spatial frequency fearful faces. Cortex. pii: S0010-9452(17)30205-8.Google Scholar
Azmitia, E. (2010). Evolution of serotonin: sunlight to suicide. In: Muller, C., Jacobs, B. (Eds.), Handbook of Behavioral Neurobiology of Serotonin. San Diego, CA: Elsevier/Academic Press, pp. 125–61.Google Scholar
Bai, Y. J., et al. (2009). Orexin A attenuates unconditioned sexual motivation in male rats. Pharmacol Biochem Behav. 91(4): 581–9.Google Scholar
Baker, J. L., et al. (2016). Robust modulation of arousal regulation, performance, and frontostriatal activity through central thalamic deep brain stimulation in healthy nonhuman primates. J Neurophysiol. 116(5): 2383–404.Google Scholar
Barabási, A. L. (2009). Scale-free networks: a decade and beyond. Science. 325: 412–13.Google Scholar
Barabasi, A. L. (2002). Linked: The New Science of Networks. Cambridge, MA: Perseus.Google Scholar
Barabasi, A.-L. (2015). Spectrum of controlling and observing complex networks. Nat Phys. 11: 779–96.Google Scholar
Barabasi, A. L., Albert, R. (1999). Emergence of scaling in random networks. Science. 286(5439): 509–12.CrossRefGoogle ScholarPubMed
Barabási, A. L. et al. (2011). Network medicine: a network-based approach to human disease. Nat Rev Genet. 12(1): 5668.Google Scholar
Bargiello, T. A., Young, M. W. (1984). Molecular genetics of a biological clock in Drosophila. Proc Natl Acad Sci U S A. 81(7): 2142–6.Google Scholar
Barrett, L. F. (2017). How Emotions Are Made. Boston, MA: Houghton Mifflin Harcourt.Google Scholar
Bartolomeo, P., Chokron, S. (2002). Orienting of attention in left unilateral neglect. Neurosci Biobehav Rev. 26(2): 217–34.Google Scholar
Barzel, B., Liu, Y., Barabasi, A.-L. (2015). Constructing minimal models for complex system dynamics. Nat Commun. 6: 18.Google Scholar
Bass, J., Lazar, M. A. (2016). Circadian time signatures of fitness and disease. Science. 354: 994–9.Google Scholar
Bassett, D. S., et al. (2006). Adaptive reconfiguration of fractal small world human brain functional networks. PNAS. 103: 19518–23.Google Scholar
Bassett, D. S., Sporns, O. (2017) Article I. Network nueroscience. Nat Neurosci. 20(3): 353–64. doi: 10.1038/nn.4502.Google Scholar
Bassetti, C. L. (2011). Sleep and stroke. In: Kryger, M. H., Roth, T., Dement, W. C. (Eds.), Principles and Practice of Sleep Medicine, 5th edn. Philadelphia, PA: Saunders, pp. 9931015.Google Scholar
Batini, C., Moruzzi, G., Palestini, M., Rossi, G. F., Zanchetti, A. (1959). Effects of complete pontine transections on the sleep-wakefulness rhythm: the midpontine pretri-geminal preparation. Arch Ital Biol. 97: 112.Google Scholar
Becker, J., et al. (Eds.). (2008). Sex Differences in the Brain. New York, NY: Oxford University Press.Google Scholar
Beckstead, R. M., Domesick, V. B., Nauta, W. J. (1979). Efferent connections of the substantia nigra and ventral tegmental area in the rat. Brain Res. 175: 191217.Google Scholar
Bedrosian, T. A., Nelson, R. J. (2014). Nitric oxide and serotonin interactions in aggression. Curr Top Behav Neurosci. 17: 131–42.Google ScholarPubMed
Behn, C., Brown, E., Scammell, T., Kopell, N. (2007). Mathematical model of network dynamics governing mouse sleep-wake behavior. J Neurophysiol. 97:382803840.Google Scholar
Beitz, A. J. (1982). The organization of afferent projections to the midbrain periaqueductal gray of the rat. Neuroscience. 7(1): 133–59.Google Scholar
Benarroch, E. E. (2013). Pedunculopontine nucleus: functional organization and clinical implications. Neurology. 80(12): 1148–55.Google Scholar
Beretzner, F., Brownstone, R. M. (2013). Lhx3–Chx10 reticulospinal neurons in locomotor circuits. J Neurosci. 33: 14681–92.Google Scholar
Berntson, G. G., Shafi, R., Sarter, M. (2002). Specific contributions of the basal forebrain corticopetal cholinergic system to electroencephalographic activity and sleep/waking behaviour. Eur J Neurosci. 16: 2453–61.Google Scholar
Berridge, C. W. (2008). Noradrenergic modulation of arousal. Brain Res Rev. 58: 117.Google Scholar
Berridge, C. W., Foote, S. L. (1991). Effects of locus coeruleus activation on EEG activity in neocortex and hippocampus. J Neurosci. 11: 3135–45.Google Scholar
Berridge, C. W., Foote, S. L. (1996). Enhancement of behavioral and electroencephalographic indices of waking following stimulation of noradrenergic beta-receptors within the medial septal region of the basal forebrain. J Neurosci. 16: 69997009.Google Scholar
Berridge, C. W., Waterhouse, B. D. (2003). The locus coeruleus-noradrenergic system: modulation of behavioral state and state-dependent cognitive processes. Brain Res Rev. 42: 3384.Google Scholar
Bharos, T. B., Kuypers, H. G., Lemon, R. N., Muir, R. B. (1981). Divergent collaterals from deep cerebellar neurons to thalamus and tectum, and to medulla oblongata and spinal cord: retrograde fluorescent and electrophysiological studies. Exp Brain Res. 42(3-4): 399410.Google Scholar
Blair, R.J., Peschardt, K.S., Budhani, S., Mitchell, D.G., Pine, D.S. (2006). Article I. The development of psychopathy. J Child Psychol Psychiatry. 47(3–4): 262–76.Google Scholar
Blanco-Centurion, C., Xu, M.C., Murillo-Rodriguez, E., et al. (2006). Adenosine and sleep homeostasis in the basal forebrain. J Neurosci. 26: 8092–100.Google Scholar
Blanco-Centurion, C., Gerashchenko, D., Shiromani, P. J. (2007). Effects of saporin-induced lesions of three arousal populations on daily levels of sleep and wake. J Neurosci. 27(51): 14041–8.Google Scholar
Blaustein, J. D., Turcotte, J. C. (1989). Estradiol-induced progestin receptor immunoreactivity is found only in estrogen receptor-immunoreactive cells in guinea pig brain. Neuroendocrinology. 49(5): 454–61.Google Scholar
Block, N. (2007). Consciousness, Function and Representation. Cambridge, MA: MIT Press.Google Scholar
Blumbergs, P. C., Jones, N. R., North, J. B. (1989). Diffuse axonal injury in head trauma. J Neurol Neurosurg Psychiatry. 52: 838–41.Google Scholar
Blumenfeld, H., McCormick, D. A. (2000). Corticothalamic inputs control the pattern of activity generated in thalamocortical networks. J Neurosci. 20(13): 5153–62.Google Scholar
Bodart, O., et al. (2017). Measures of metabolism and complexity in the brain of patients with disorders of consciousness. Neuroimage Clin. 14: 354–62.Google Scholar
Bonnelle, V., et al. (2012). Salience network integrity predicts default mode network function after traumatic brain injury. PNAS. 109: 4690–5.Google Scholar
Borbely, A. A. (1982). A two process model of sleep regulation. Human Neurobiol. 1: 195204.Google Scholar
Borbely, A. A., Tobler, I. (1985). Homeostatic and circadian principles in sleep regulation in the rat. In: McGinty, D. J., et al. (Eds.), Brain Mechanisms of Sleep. New York, NY: Raven Press, pp. 3544.Google Scholar
Boutrel, B., Cannella, N., de Lecea, L. (2010). The role of hypocretin in driving arousal and goal-oriented behaviors. Brain Res. 1324: 103–11.Google Scholar
Bremer, F. (1935). Cerveau isole et physiologie du sommeil. CR Soc Biol (Paris). 118: 1235–42.Google Scholar
Brink, E. E., Pfaff, D. W. (1981). Supraspinal and segmental input to lumbar epaxial motoneurons in the rat. Brain Res. 226: 4360.Google Scholar
Broadbent, D. E. (1971). Decision and Stress. London: Academic Press.Google Scholar
Brookes, M. J., et al. (2011). Investigating the electrophysiological basis of resting state networks using magnetoencephalography. PNAS. 108: 16783–8.Google Scholar
Broom, L., et al. (2017). A translational approach to capture gait signatures of neurological disorders in mice and humans. Sci Rep. 7(1): 3225. doi: 10.1038/s41598-017-03336-1.Google Scholar
Brown, R. E., McKenna, J. T. (2015). Turning a negative into a positive: ascending GABAergic control of cortical activation and arousal. Front Neurol. 6: 135. doi: 10.3389/fneur.2015.00135.Google Scholar
Brown, R. E., et al. (2006). Electrophysiological characterization of neurons in the dorsolateral pontine rapid-eye-movement sleep induction zone of the rat: intrinsic membrane properties and responses to carbachol and orexins. Neuroscience. 143(3): 739–55.Google Scholar
Brown, E. N., Lydic, R., Schiff, N. D. (2010). General anesthesia, sleep, and coma. New Engl J Med. 363(27): 2638–50.Google Scholar
Brust, J., (2000). In Principles of Neural Science (4th edition). New York, NY: McGraw-Hill, pp. 1302–16.Google Scholar
Buchanan, G., Richerson, G. (2010). Central serotonin neurons are required for arousal to CO2. Proc Natl Acad Sci. 107: 16354–9.Google Scholar
Buckner, R. L., et al. (2008). The brain's default network anatomy, function, and relevance to disease. Ann NY Acad Sci. 1124: 138.Google Scholar
Buckner, R. L., (2009). Cortical hubs revealed by intrinsic functional connectivity: mapping, assessment of stability, and relation to Alzheimer's disease. J Neurosci. 29(6): 1860–73.Google Scholar
Buckner, R. L., (2011). The organization of the human cerebellum estimated by intrinsic functional connectivity. J Neurophysiol. 106: 2322–45.Google Scholar
Bullmore, E., Sporns, O. (2009). Complex brain networks: graph theoretical analysis of structural and functional systems. Nature Rev Neurosci. 10(3): 186–98.Google Scholar
Butler, M. P., Karatsoreos, I. N., LeSauter, J., Silver, R. (2012). Dose-dependent effects of androgens on the circadian timing system and its response to light. Endocrinology. 153(5): 2344–52.Google Scholar
Buzsaki, G. (2006). Rhythms of the Brain. NewYork, NY: Oxford University Press.Google Scholar
Buzsáki, G., Schomburg, E. W. (2015). What does gamma coherence tell us about inter-regional neural communication? Nat Neurosci. 18(4): 484–9.CrossRefGoogle ScholarPubMed
Buzsaki, G., Bickford, R. G., Ponomareff, G., et al. (1998). Nucleus basalis and thalamic control of neocortical activity in the freely moving rat. J Neurosci. 8: 4007–26.Google Scholar
Buzsáki, G., Geisler, C., Henze, D. A., Wang, X. J. (2004). Interneuron diversity series: circuit complexity and axon wiring economy of cortical interneurons. Trends Neurosci. 27: 186–93.Google Scholar
Caballero, P. E. J. (2010). Bilateral paramedian thalamic artery infarcts: report of 10 cases. J Stroke Cardiovasc Dis. 19: 283–9.Google Scholar
Caggiano, V., et al. (2017). Midbrain circuits that set locomotor speed and gait selection. Nature. 553(7689): 455–60.Google Scholar
Caldarelli, G. (2007). Scale-Free Networks. Oxford: Oxford University Press.Google Scholar
Calderon, D. P., Proekt, A., Pfaff, D. (2018). Activation of large neurons in the medullary reticular formation regulates cortical and behavioral arousal. Nat Neurosci. (submitted).Google Scholar
Calderon, D. P., et al. (2016). Generalized CNS arousal: an elementary force within the vertebrate nervous system. Neurosci Biobehav Rev. 68: 167–76.Google Scholar
Caldwell, H., Albers, H. E. (2016). Oxytocin, vasopressin, and the motivational forces that drive social behaviors. Curr Topics Behav Neurosci. 27: 51103.Google Scholar
Cantero, J. L., Atienza, M. (2005). The role of neural synchronization in the emergence of cognition across the wake-sleep cycle. Rev Neurosci. 16(1): 6983.Google Scholar
Cape, E. G., Jones, B. E. (2000). Effects of glutamate agonist versus procaine microinjections into the basal forebrain cholinergic cell area upon gamma and theta EEG activity and sleep–wake state. Eur J Neurosci. 12(6): 2166–84.Google Scholar
Capelli, P., et al. (2017). Locomotor speed control circuits in the caudal brainstem. Nature. 551(7680): 373–7.CrossRefGoogle ScholarPubMed
Carroll, M. E., Anker, J. J. (2010). Sex differences and ovarian hormones in animal models of drug dependence. Horm Behav. 58(1): 4456.Google Scholar
Carter, M. E., Adamantidis, A., Ohtsu, H., Deisseroth, K., de Lecea, L. (2009a). Sleep homeostasis modulates hypocretin-mediated sleep-to-wake transitions. J Neurosci. 29(35): 10939–49.Google Scholar
Carter, M. E., Borg, J. S., de Lecea, L. (2009b). The brain hypocretins and their receptors: mediators of allostatic arousal. Curr Opin Pharmacol. 9(1): 3945. Review.Google Scholar
Carter, M. E., et al. (2010). Tuning arousal with optogenetic modulation of locus coeruleus neurons. Nat Neurosci. 13(12): 1526–33.Google Scholar
Carter, M. E., Brill, J., Bonnavion, P., et al. (2012). Mechanism for hypocretin-mediated sleep-to-wake transitions. Proc Natl Acad Sci U S A. 109: E2635–44.Google Scholar
Carter, M. E., de Lecea, L., Adamantidis, A. (2013). Functional wiring of hypocretin and LC-NE neurons: implications for arousal. Front Behav Neurosci. 20(7): 43.Google Scholar
Casaratto, S. (2016). Stratification of unresponsive patients by an independently validated index of brain complexity. Ann Neurol. 80: 718–29.Google Scholar
Cedarbaum, J. M., Aghajanian, G. K. (1978). Afferent projections to the rat locus coeruleus as determined by a retrograde tracing technique. J Comp Neurol. 178(1): 116.Google Scholar
Cembrowski, M. S., et al. (2016). Hipposeq: a comprehensive RNA-seq database of gene expression in hippocampal principal neurons. Elife. 5: e14997.Google Scholar
Chamberlin, N. L., Saper, C. B. (1992). Topographic organization of cardiovascular responses to electrical and glutamate microstimulation of the parabrachial nucleus in the rat. J Comp Neurol. 326(2): 245–62.Google Scholar
Chamberlin, N. L., Saper, C. B. (1994). Topographic organization of respiratory responses to glutamate microstimulation of the parabrachial nucleus in the rat. J Neurosci. 14: 6500–10.CrossRefGoogle ScholarPubMed
Chandler, D. J., Waterhouse, B. D., Gao, W. J. (2014). New perspectives on catecholaminergic regulation of executive circuits: evidence for independent modulation of prefrontal functions by midbrain dopaminergic and noradrenergic neurons. Front Neural Circuits. 8: 53. doi: 10.3389/fncir.2014.00053.Google Scholar
Chemelli, R., et al. (1999). Narcolepsy in orexin knockout mice: molecular genetic and sleep regulation. Cell. 98: 437–51.Google Scholar
Chen, J., Randeva, H. S. (2004). Genomic organization of mouse orexin receptors. Mol. Endocrinol. 18: 2790–804.Google Scholar
Chen, C. T., Dun, S. L., Kwok, E. H., Dun, N. J., Chang, J. K. (1999). Orexin A-like immunoreactivity in the rat brain. Neurosci Lett. 260: 161–4.Google Scholar
Chen, C., et al. (2015). Testosterone modulates preattentive sensory processing and involuntary attention switches to emotional voices. J Neurophysiol. 113(6): 1842–9.Google Scholar
Chennu, S., et al. (2017). Brain networks predict metabolism, diagnosis and prognosis at the bedside in disorders of consciousness. Brain. 140: 2120–32.Google Scholar
Chou, T. C., Bjorkum, A. A., Gaus, S. E., et al. (2002). Afferents to the ventrolateral preoptic nucleus. J Neurosci. 22: 977–90.Google Scholar
Chou, T. C., Scammell, T. E., Gooley, J. J., et al. (2003). Critical role of dorsomedial hypothalamic nucleus in a wide range of behavioral circadian rhythms. J Neurosci. 23(33): 10691–702.Google Scholar
Christensen, J. A., et al. (2015). Sleep-stage transitions during polysomnographic recordings as diagnostic features of type 1 narcolepsy. Sleep Med. 16(12): 1558–66.Google Scholar
Chung, S., et al. (2017). Identification of preoptic sleep neurons using retrograde labeling and gene profiling. Nature. 545: 477–82.Google Scholar
Ciriello, J., de Oliveira, C. V. R., Masoumeh M., et al. (2002). Estrogen alters the cardiovascular responses to activation of rostral ventrolateral medulla in the female. Soc Neurosci Abs. 768: 6.Google Scholar
Claassen, J., et al. (2013). Recommendations on the use of EEG monitoring in critically ill patients: consensus statement from the neurointensive care section of the ESICM. Intensive Care Med. 39(8): 1337–51.Google Scholar
Claassen, J., (2016). Bedside quantitative electroencephalography improves assessment of consciousness in comatose subarachnoid hemorrhage patients. Ann Neurol. 80(4): 541–53.Google Scholar
Claiborne, J. A., Nag, S., Mokha, S. S. (2009). Article I. Estrogen-dependent, sex-specific modulation of mustard oil-induced secondary thermal hyperalgesia by orphanin FQ in the rat. Neurosci Lett. 456(2): 5963. doi: 10.1016/j.neulet.2009.03.106. Epub 2009 Apr 5.Google Scholar
Clipperton-Allen, A. E., et al. (2010). Agonistic behavior in males and females: Effects of an estrogen receptor beta agonist in gonadectomized and gonadally intact mice. Psychoneuroendocrinology. 35(7): 1008–22.Google Scholar
Cohen, J. E. (1995). Unexpected dominance of high frequencies in chaotic nonlinear population models. Nature. 378: 610–16.Google Scholar
Cohen, M. S., Schwartz-Giblin, S., Pfaff, D. W. (1987). Brainstem reticular stimulation facilitates back muscle motoneuronal responses to pudendal nerve input. Brain Res. 405: 155–61.Google Scholar
Colvin, G. B., Whitmoyer, D. I., Lisk, R. D. et al. (1968). Changes in sleep-wakefulness in female rats during circadian and estrous cycles. Brain Res. 7(2): 173–81.Google Scholar
Colvin, G. B., Whitmoyer, D. I., Sawyer, C. H. (1968). Circadian sleep-wakefulness patterns in rats after ovariectomy and treatment with estrogen. Exp Neurol. 25(4): 616–25.Google Scholar
Conrad, L. A., Pfaff, D. W. (1975). Axonal projections of medial preoptic and anterior hypothalamic neurons. Science. 190: 1112–14.Google Scholar
Conrad, L., Leonard, C., Pfaff, D. (1974). Connections of the median and dorsal raphe nuclei in the rat: an autoradiographic and degeneration study. J Comp Neurol. 156: 179205.Google Scholar
Conrad, L. C. A., Pfaff, D. W. (1976a). Efferents from medial basal forebrain and hypothalamus in the rat. I. An autoradiography study of the medial preoptic area. J. Comp. Neurol., 169: 185220.Google Scholar
Conrad, L. C. A., Pfaff, D. W. (1976b). Efferents from medial basal forebrain and hypothalamus in the rat. II. An autoradiography study of the anterior hypothalamus. J. Comp. Neurol., 169: 221–62.Google Scholar
Cools, R., Nakamura, K., Daw, N. (2011). Serotonin and dopamine: unifying affective, activational and decision functions. Neuropsychopharmacology. 36: 98113.Google Scholar
Corazzol, M., et al. (2017). Restoring consciousness with vagal nerve stimulation. Curr Biol. 27, R1R3, September 25, 2017 © 2017Elsevier Ltd. R1Google Scholar
Cottingham, S. L., Femano, P. A., Pfaff, D. W. (1987). Electrical stimulation of the midbrain central gray facilitates reticulospinal activation of axial muscle EMG. Exp Neurol. 97(3): 704–24.Google Scholar
Cottingham, S. L., Femano, P. A., Pfaff, D. W. (1988). Vestibulospinal and reticulospinal interactions in the activation of back muscle EMG in the rat. Exp Brain Res. 73(1): 198208.Google Scholar
Courtoy, P. J., Boyles, J. (1983). Fibronectin in the microvasculature: localization in the pericyte-endothelial interstitium. J Ultrastruct Res. 83: 258–73.Google Scholar
Crick, F., Koch, C. (2003). A framework for consciousness. Nat Neurosci. 6: 119–29.Google Scholar
Csete, M., Doyle, J. (2004). Bowties, metabolism and disease. Trends Biotechnol. 22: 446–50.Google Scholar
Cullinan, W. E., Zaborszky, L. (1991). Organization of ascending hypothalamic projections to the rostral forebrain with special reference to the innervation of cholinergic projection neurons. J Comp Neurol. 306: 631–67.Google Scholar
Curtis, A. L., Valentino, R. J. (1994). Corticotropin-releasing factor neurotransmission in locus coeruleus: a possible site of antidepressant action. Brain Res Bull. 35(5–6): 581–7.Google Scholar
Curtis, A. L., Bello, N. T., Connolly, K. R., Valentino, R. J. (2002). Corticotropin-releasing factor neurones of the central nucleus of the amygdala mediate locus coeruleus activation by cardiovascular stress. J Neuroendocrinol. 14(8): 667–82.Google Scholar
Da Silva, J. A., et al. (2018). Dopamine neuron activity before initiation gates and invigorates future movements. Proc Natl Acad Sci. 554: 244–50.Google Scholar
Danielle, S., Bassett, B. S., Olaf Sporns, O. (2017). Network neuroscience. Nat Neurosci. 20: 353–64.Google Scholar
Datta, S., Hobson, J. A. (1995). Suppression of ponto-geniculo-occipital waves by neurotoxic lesions of pontine caudo-lateral peribrachial cells. Neuroscience. 67(3): 703–12.Google Scholar
Datta, S., Maclean, R. R. (2007). Neurobiological mechanisms for the regulation of mammalian sleep-wake behavior: reinterpretation of historical evidence and inclusion of contemporary cellular and molecular evidence. Neurosci Biobehav Rev. 31: 775824.Google Scholar
Datta, S., Siwek, D. F. (2002). Single cell activity patterns of pedunculopontine tegmentum neurons across the sleep-wake cycle in the freely moving rats. J Neurosci Res. 70: 611–21.Google Scholar
Davidson, R., Begley, S. (2012). The Emotional Life of your Brain. NewYork, NY: Hudson St. Press (Penguin).Google Scholar
Davis, G. E., Senger, D. R. (2005). Endothelial extracellular matrix: biosynthesis, remodeling, and functions during vascular morphogenesis and neovessel stabilization. Circ Res. 97: 1093–107.Google Scholar
de Biase, S., et al. (2017). Investigational therapies for the treatment of narcolepsy. Expert Opin Investigat Drugs. 26(8): 953–63.Google Scholar
Deco, G., Hagmann, P., Hudetz, A. G., Tononi, G. (2014). Modeling resting-state functional networks when the cortex falls asleep: local and global changes. Article IV. Cereb Cortex. 24(12): 3180–94. doi: 10.1093/cercor/bht176. Epub 2013 Jul 10.Google Scholar
de Falco, F. A, et al. (1994). Bilateral thalamic damage, cortical hypometabolism and behavioural disturbances. Eur J Neurol. 1(2): 165–9.Google Scholar
de Lecea, L., Kilduff, T. S., Peyron, C., et al. (1998). The hypocretins: hypothalamus-specific peptides with neuroexcitatory activity. Proc. Natl Acad Sci. 95: 322–7.Google Scholar
Dehaene, S., Christen, Y. (2011). Characterizing Consciousness: From Cognition to the Clinic? Heidelberg: Fondation IPSEN, pp. 5583.Google Scholar
Dehaene, S., et al. (2001a). Cerebral mechanisms of word masking and unconscious repetition priming. Nat Neurosci. 4: 752–8.Google Scholar
Dehaene, S., (2001b). Toward a cognitive neuroscience of consciousness. Cognition. 79: 137.Google Scholar
Dehaene, S., Lau, H., Kouider, S. (2017). What is consciousness and could machines have it? Science. 358: 486–92.Google Scholar
Del Negro, C. A., Funk, G. D., Feldman, J. L. (2018). Breathing matters. Nat Rev Neurosci. 19(6): 351367.Google Scholar
DeLuca, D. S., Levin, J. Z., Sivachenko, A., et al. (2012). RNA-SeQC: RNA-seq metrics for quality control and process optimization. Bioinformatics. 28, 1530–2.Google Scholar
Demertzi, A., et al. (2015). Intrinsic functional connectivity differentiates minimally conscious from unresponsive patients. Brain. 138(Pt 9): 2619–31.Google Scholar
Dempsey, E. W., Morison, R. S. (1942a). The production of rhythmically recurrent cortical potentials after localized thalamic stimulation. Am J Physiol. 135: 293300.Google Scholar
Dempsey, E. W., Morison, R. S. (1942b). The interaction of certain spontaneous and induced cortical potentials. Am J Physiol. 135: 301–8.Google Scholar
Dennett, D. C. (1991). Consciousness Explained. Boston, MA: Little Brown.Google Scholar
Denno, D. W. (1990). Biology and Violence, from Birth to Adulthood. Cambridge: Cambridge University Press.Google Scholar
Denoyer, M., Sallanon, M., Buda, C., Kitahama, K., Jouvet, M. (1991). Neurotoxic lesion of the mesencephalic reticular formation and/or the posterior hypothalamus does not alter waking in the cat. Brain Res. 539: 287303.Google Scholar
Deurveilher, S., Semba, K. (2004). Indirect projections from the suprachiasmatic nucleus to major arousal-promoting cell groups in rat: implications for the circadian control of behavioral state. Neuroscience. 130: 165–84.Google Scholar
Devi, L., Fricker, L. (2016a). Transmitters and peptides: basic principles. In: Pfaff, D. W., Volkow, N. D. (Eds.), Neuroscience in the 21st Century (2nd edition, volume 3). New York, NY: Springer Verlag, pp. 1746–62.Google Scholar
Devi, L., Fricker, L. (2016b). Transmitter and peptide receptors: basic principles. In: Pfaff, D. W., Volkow, N. D. (Eds.), Neuroscience in the 21st Century (2nd edition, volume 3). New York, NY: Springer Verlag, pp. 1763–86.Google Scholar
Devidze, N., et al. (2008). Presynaptic actions of opioid receptor agonists in ventromedial hypothalamic neurons in estrogen-and oil-treated female mice. Neuroscience. 152(4): 942–9.Google Scholar
Devidze, N., (2010). Estradiol regulation of lipocalin-type prostaglandin D synthase promoter activity: evidence for direct and indirect mechanisms. Neuroscience Lett. 474(1): 1721.Google Scholar
Dietrich, T., et al. (2001). Effects of blood estrogen level on cortical activation patterns during cognitive activation as measured by functional MRI. Neuroimage. 13(3): 425–32.Google Scholar
Dijk, D.J., Winsky-Sommerer, R. (2016). Sleep and dreamless mice. Nature. 539: 364–5.Google Scholar
Dina, O. A., et al. (2001). Sex hormones regulate the contribution of PKCepsilon and PKA signalling in inflammatory pain in the rat. Eur J Neurosci. 13(12): 2227–33.Google Scholar
Divac, I., Björklund, A., Lindvall, O., Passingham, R. E. (1978). Converging projections from the mediodorsal thalamic nucleus and mesencephalic dopaminergic neurons to the neocortex in three species. J Comp Neurol. 180(1): 5971.Google Scholar
Dobin, A., Davis, C. A., Schlesinger, F., et al. (2013). STAR: ultrafast universal RNA-seq aligner. Bioinformatics. 29: 1521.Google Scholar
Doesburg, S. M., Roggeveen, A. B., Kitajo, K., Ward, L. M. (2008). Large-scale gamma-band phase synchronization and selective attention. Cerebr Cortex. 18(2): 386–96.Google Scholar
Dordea, A. C., et al. (2016). Androgen-sensitive hypertension associated with soluble guanylate cyclase-α1 deficiency is mediated by 20-HETE. Am J Physiol Heart Circ Physiol. 310(11): H1790–800.Google Scholar
Doria, V. et al. (2010). Emergence of resting state networks in the preterm human brain. PNAS. 107: 20015–20.Google Scholar
Doyle, J., Csete, M. (2005). Motifs, stability and control. PLoS Biol. 3: e392.Google Scholar
Drew, T., Rossignol, S. (1990a). Functional organization within the medullary reticular formation of intact unanesthetized cat. I. Movements evoked by microstimulation. J Neurophysiol. 64(3): 767–81.Google Scholar
Drew, T., Rossignol, S. (1990b). Functional organization within the medullary reticular formation of intact unanesthetized cat. II. Electromyographic activity evoked by microstimulation. J Neurophysiol. 64(3): 782–95.Google Scholar
Drew, T., Dubuc, R., Rossignol, S. (1986). Discharge patterns of reticulospinal and other reticular neurons in chronic, unrestrained cats walking on a treadmill. J Neurophysiol. 55(2): 375401.Google Scholar
Dringenberg, H. C., Olmstead, M. C. (2003). Integrated contributions of basal forebrain and thalamus to neocortical activation elicited by pedunculopontine tegmental stimulation in urethane-anesthetized rats. Neuroscience. 119: 839–53.Google Scholar
Drover, J. D., Schiff, N. D., Victor, J. D. (2010). Dynamics of coupled thalamocortical modules. J. Comput Neurosci. 28: 605–16.Google Scholar
Duffy, E. (1962). Activation and Behavior. New York, NY: Wiley.Google Scholar
Dupré, C., et al. (2010). Histaminergic responses by hypothalamic neurons that regulate lordosis and their modulation by estradiol. Proc Natl Acad Sci U S A. 107(27): 12311–6.Google Scholar
Easton, A., Dwyer, E., Pfaff, D. W. (2006). Estradiol and orexin-2 saporin actions on multiple forms of behavioral arousal in female mice. Behav Neurosci. 120(1): 19.Google Scholar
Eberhart, J. A., Morrell, J. I., Krieger, M. S., Pfaff, D. W. (1985). An autoradiographic study of projections ascending from the midbrain central gray, and from the region lateral to it, in the rat. J Comp Neurol. 241(3): 285310.Google Scholar
Edlow, B. L., et al. (2017). Early detection of consciousness in patients with acute severe traumatic brain injury. Brain. 140(9): 2399–414.Google Scholar
Edwards, S. B., de Olmos, J. S. (1976). Autoradiographic studies of the projections of the midbrain reticular formation: ascending projections of nucleus cuneiformis. J Comp Neurol. 165(4): 417–31.Google Scholar
Ekstrand, M. I., Nectow, A. R., Knight, Z. A., et al. (2014). Molecular profiling of neurons based on connectivity. Cell. 157, 1230–42.Google Scholar
Elam, M., Svensson, T. H., Thoren, P. (1985). Differentiated cardiovascular afferent regulation of locus coeruleus neurons and sympathetic nerves. Brain Res. 358(1–2): 7784.Google Scholar
Elam, M., Thorén, P., Svensson, T. H. (1986). Locus coeruleus neurons and sympathetic nerves: activation by visceral afferents. Brain Res. 375(1): 117–25.Google Scholar
Elam, M., Yao, T., Svensson, T. H., Thoren, P. (1984). Regulation of locus coeruleus neurons and splanchnic, sympathetic nerves by cardiovascular afferents. Brain Res. 290(2): 281–7.Google Scholar
Elisevich, K. V., Hrycyshyn, A. W., Flumerfelt, B. A. (1985). Cerebellar, medullary and spinal afferent connections of the paramedian reticular nucleus in the cat. Brain Res. 332(2): 267–82.Google Scholar
Elmquist, J. K., et al. (1997). Leptin activates neurons in ventrobasal hypothalamus and brainstem. Endocrinology. 138(2): 839–42.Google Scholar
Ericson, H., Blomqvist, A. (1988). Tracing of neuronal connections with cholera toxin subunit B: light and electron microscopic immunohistochemistry using monoclonal antibodies. J Neurosci Meth. 24: 225–35.Google Scholar
España, R. A., Reis, K. M., Valentino, R. J., Berridge, C. W. (2005). Organization of hypocretin/orexin efferents to locus coeruleus and basal forebrain arousal-related structures. J Comp Neurol. 481(2): 160–78.Google Scholar
Evans, S. M., Foltin, R. W. (2010). Does the response to cocaine differ as a function of sex or hormonal status in human and non-human primates? Horm Behav. 58(1): 1321.Google Scholar
Eysenck, H. J., Eysenck, S. B. (1967). On the unitary nature of extraversion. Acta Psychol (Amst). 26(4): 383–90.Google Scholar
Faber, D. S., et al. (1989). Neuronal networks underlying the escape response in goldfish. Ann N Y Acad Sci. 563: 1133.Google Scholar
Fadok, J. P., et al. (2017). A competitive inhibitory circuit for selection of active and passive fear responses. Nature. 542: 96105.Google Scholar
Fantin, A., Maden, C. H., Ruhrberg, C. (2009). Neuropilin ligands in vascular and neuronal patterning. Biochem Soc Trans. 37(Pt 6): 1228–32.Google Scholar
Fardin, V., Oliveras, J. L., Besson, J. M. (1984). Projections from the periaqueductal gray matter to the B3 cellular area (nucleus raphe magnus and nucleus reticularis paragigantocellularis) as revealed by the retrograde transport of horseradish peroxidase in the rat. J Comp Neurol. 223(4): 483500.Google Scholar
Feliers, D., Chen, X., Akis, N., et al. (2005). VEGF regulation of endothelial nitric oxide synthase in glomerular endothelial cells. Kidney Int. 68: 1648–59.Google Scholar
Fetcho, J. R., McLean, D. L. (2010). Some principles of organization of spinal neurons underlying locomotion in zebrafish and their implications. Annals N Y Acad Sci. 1198: 94104.Google Scholar
Filosa, J. A., Iddings, J. A. (2013). Astrocyte regulation of cerebral vascular tone. Am J Physiol Heart Circ Physiol. 305: H609619.Google Scholar
Fins, J. J. (2015). Brain Comes to Mind. Cambridge: Cambridge University Press.Google Scholar
Fiset, P., et al. (1999). Brain mechanisms of propofol-induced loss of consciousness in humans: a positron emission tomographic study. J Neurosci. 19(13): 5506–13.Google Scholar
Flavell, S. W., et al. (2013). Serotonin and the neuropeptide PDF initiate and extend opposing behavioral states in C. elegans. Cell. 154(5): 1023–35.Google Scholar
Fogerson, P. M., Huguenard, J. R. (2016). Tapping the brakes: cellular and synaptic mechanisms that regulate thalamic oscillations. Neuron. 92(4): 687704.Google Scholar
Fontani, G., et al. (2004). Attentional, emotional and hormonal data in subjects of different ages. Eur J Appl Physiol. 92: 452–61.Google Scholar
Forgacs, P. B., et al. (2017). Dynamic regimes of neocortical activity linked to corticothalamic integrity correlate with outcomes in acute anoxic brain injury after cardiac arrest. Ann Clin Transl Neurol. 4(2): 119–29.Google Scholar
Fornito, A., et al. (2012). Competitive and cooperative dynamics of large-scale brain functional networks supporting recollection. PNAS. 109: 12788–93.Google Scholar
Fox, M. D., et al. (2005). The human brain is intrinsically organized into dynamic anticorrelated functional networks. PNAS. 102: 9673–8.Google Scholar
Fox, M. D., (2006). Spontaneous neuronal activity distinguishes human dorsal and ventral attention systems. PNAS. 103: 10046–51.Google Scholar
Fregosi, M., et al. (2017). Corticobulbar projections from distinct motor cortical areas to the reticular formation in macaque monkeys. Eur J Neurosci. doi: 10.1111/ejn.13576.Google Scholar
Fridman, E. A., et al. (2014). Regional cerebral metabolic patterns demonstrate the role of anterior forebrain mesocircuit dysfunction in the severely injured brain. Proc Natl Acad Sci U S A. 111(17): 6473–8.Google Scholar
Frohlich, J., et al. (2001). Statistical analysis of measures of arousal in ovariectomized female mice. Horm Behav. 39(1): 3947.Google Scholar
Frohlich, J., (2002). Statistical analysis of hormonal influences on arousal measures in ovariectomized female mice. Horm Behav. 2(4): 414–23.Google Scholar
Frohmader, K. S., Pitchers, K. K., Balfour, M. E., Coolen, L. M. (2010). Mixing pleasures: review of the effects of drugs on sex behavior in humans and animal models. Horm Behav. 58(1): 149–62.Google Scholar
Fuller, P. M., Gooley, J. J., Saper, C. B. (2006). Neurobiology of the sleep–wake cycle: sleep architecture, circadian regulation, and regulatory feedback. J Biol Rhythms. 21: 482–93.Google Scholar
Fuller, P. M., Saper, C. B., Lu, J. (2007). The pontine REM switch: past and present. J Physiol. 584: 735–41.Google Scholar
Fuller, P. M., et al. (2011). Reassessment of the structural basis of the ascending arousal system. J Comp Neurol. 519(5): 933–56.Google Scholar
Fulwiler, C. E., Saper, C. B. (1984). Subnuclear organization of the efferent connections of the parabrachial nucleus in the rat. Brain Res. 319: 229–59.Google Scholar
Funato, H., et al. (2016). Forward-genetics analysis of sleep in randomly mutagenized mice. Nature. 539: 378–87.Google Scholar
Gagnidze, K., Weil, Z., Khattak, M., Pfaff, D. (2010). Estrogen-induced chromatin remodeling and gene transcription in ventromedial hypothalamus. Soc Neurosci. Abstract #495.11 (Poster)Google Scholar
Gallager, D. W., Pert, A. (1978). Afferents to brain stem nuclei (brain stem raphe, nucleus reticularis pontis caudalis and nucleus gigantocellularis) in the rat as demonstrated by microiontophoretically applied horseradish peroxidase. Brain Res. 144(2): 257–75.Google Scholar
Gao, J., Barzel, B., Barabasi, A.-L. (2016). Universal resilience patterns in complex systems. Nature. 530: 307–12.Google Scholar
Garcia-Cardena, G., Fan, R., Shah, V., et al. (1998). Dynamic activation of endothelial nitric oxide synthase by Hsp90. Nature. 392: 821–4.Google Scholar
Garcia-Rill, E., Skinner, R. D., Gilmore, S. A., Owings, R. (1983). Connections of the mesencephalic locomotor region (MLR) II. Afferents and efferents. Brain Res Bull. 10(1): 6371.Google Scholar
Garcia-Rill, E. (2015a). The physiology of the pedunculopontine nucleus: implications for deep brain stimulation. J Neural Transm (Vienna). 122(2): 225–35.Google Scholar
Garcia-Rill, E. (2015b). Waking and the Reticular Activating System in Health and Disease. San Diego, CA: Academic Press/Elsevier.Google Scholar
Garel, S., López-Bendito, G. (2014). Inputs from the thalamocortical system on axon pathfinding mechanisms. Curr Opin Neurobiol. 27: 143–50.Google Scholar
Garey, J., et al. (2002). Temporal and spatial quantitation of reproductive behaviors among mice housed in a seminatural environment. Horm Behav. 42: 294306.Google Scholar
Garey, J., (2003). Genetic contributions to generalized arousal of brain and behavior. Proc Natl Acad Sci U S A. 100(19): 11019–22.Google Scholar
Gaus, S. E., et al. (2002). Ventrolateral preoptic nucleus contains sleep-active, galaninergic neurons in multiple mammalian species. Neuroscience. 115(1): 285–94.Google Scholar
Gennarelli, T. A, et al. (1982). Diffuse axonal injury and traumatic coma in the primate. Ann Neurol. 12: 564–74.Google Scholar
Gerashchenko, D., Kohls, M. D., Greco, M., et al. (2001). Hypocretin-2-saporin lesions of the lateral hypothalamus produce narcoleptic-like sleep behavior in the rat. J Neurosci. 21: 7273–83.Google Scholar
Gezelius, H., López-Bendito, G. (2017). Thalamic neuronal specification and early circuit formation. Dev Neurobiol. 77(7): 830–43.Google Scholar
Giacino, J., Fins, J. J., Machado, A., Schiff, N. D. (2012). Central thalamic deep brain stimulation to promote recovery from chronic posttraumatic minimally conscious state: challenges and opportunities. Neuromodulation. 15(4): 339–49.Google Scholar
Giacino, J. T., Kalmar, K., Whyte, J. (2004). The JFK Coma Recovery Scale-Revised: measurement characteristics anddiagnostic utility. Arch Phys Med Rehab. 85: 2020–9.Google Scholar
Giacino, J. T., et al. (2014). Disorders of consciousness after acquired brain injury: the state of the science. Nature Rev Neurol. 10(2): 99114.Google Scholar
Gilbert, K. A., Lydic, R. (1990). Parabrachial neuron discharge in the cat is altered during the carbachol-induced REM sleep-like state (DCarb) Neurosci Lett. 120: 241–4.Google Scholar
Gillies, G. E., McArthur, S. (2010). Estrogen actions in the brain and the basis for differential action in men and women: a case for sex-specific medicines. Pharmacological Rev. 62(2): 155–98.Google Scholar
Gloor, P., Ball, G., Schaul, N. (1977). Brain lesions that produce delta waves in the EEG. Neurology. 27(4): 326–33.Google Scholar
Gogos, J., et al. (1998). Catechol-O-methyltransferase-deficient mice exhibit sexually dimorphic changes in catecholamine levels and behavior. Proc Natl Acad Sci U S A. 95: 9991–6.Google Scholar
Goldfoot, D., Baum, M. Initiation of mating behavior in developing male rats following peripheral electric shock. Physiol Behav. 8: 857–63.Google Scholar
González-Cuello, A., Milanés, M. V., Laorden, M. L. (2004). Increase of tyrosine hydroxylase levels and activity during morphine withdrawal in the heart. Eur J Pharmacol. 506(2): 119–28.Google Scholar
Gooley, J. J., Schomer, A., Saper, C. B. (2006). The dorsomedial hypothalamic nucleus is critical for the expression of food-entrainable circadian rhythms. Nat Neurosci. 9(3): 398407.Google Scholar
Graham, M. D., Pfaus, J. G. (2012). The effect of specific dopamine receptor antagonists in the medial preoptic area on the sexual behaviour of female rats. Pharmacol Biochem Behav. 102: 532–9.Google Scholar
Greco, M. A., Fuller, P. M., Jhou, T. C., et al. (2008). Opioidergic projections to sleep-active neurons in the ventrolateral preoptic nucleus. Brain Res. 1245: 96107.Google Scholar
Gregg, T. R. (2003). Cortical and limbic neural circuits mediating aggressive behavior. In: Mattson, M. P. (Ed.), Neurobiology of Aggression. Totowa, NJ: Humana Press, pp. 121.Google Scholar
Gritti, I., et al. (2006). Stereological estimates of the basal forebrain cell population in the rat, including neurons containing choline acetyltransferase, glutamic acid decarboxylase or phosphate-activated glutaminase and colocalizing vesicular glutamate transporters. Neuroscience. 143(4): 1051–64.Google Scholar
Groenewegen, H. J., Berendse, H. W. (1994). The specificity of the ‘nonspecific’ midline and intralaminar thalamic nuclei. Trends Neurosci. 17(2): 52–7.Google Scholar
Grove, E. A. (1988). Neural associations of the substantia innominata in the rat: afferent connections. J Comp Neurol. 277: 315–46.Google Scholar
Guarraci, F. A. (2010). Sex, drugs and the brain: the interaction between drugs of abuse and sexual behavior in the female rat. Horm Behav. 58(1): 138–48.Google Scholar
Guiard, B. P., et al. (2008). Functional interactions between dopamine, serotonin and norepinephrine neurons. Int J Neuropsychopharmacol. 11: 625–39.Google Scholar
Gulia, K. K., Mallick, H. N., Kumar, V. M. (2003). Orexin A (hypocretin-1) application at the medial preoptic area potentiates male sexual behavior in rats. Neuroscience. 116(4): 921–3.Google Scholar
Guyenet, P., et al. (1996). Role of medulla oblongata in generation of sympathetic and vabal outflows. In: Holstege, G., Bandler, R., Saper, C.B., (1996). The Emotional Motor System: Progress in Brain Research Vol. 107. ElsevierGoogle Scholar
Hadj-Bouziane, F., et al. (2012). Amygdala lesions disrupt modulation of functional MRI activity evoked by facial expression in the monkey inferior temporal cortex. Proc Natl Acad Sci U S A. 109(52): E3640–8.Google Scholar
Hadjimarkou, M. M., et al. (2008). Estradiol suppresses rapid eye movement sleep and activation of sleep-active neurons in the ventrolateral preoptic area. Eur J Neurosci. 27(7): 1780–92.Google Scholar
Hagemann, D., Waldstein, S. R., Thayer, J. F. (2003). Central and autonomic nervous system integration in emotion. Brain Cogn. 52(1): 7987.Google Scholar
Haglund, L., Köhler, C., Ross, S. B., Kelder, D. (1979). Forebrain projections of the ventral tegmentum as studied by axonal transport of [3H]dopamine in the rat. Neurosci Lett. 12: 301–6.Google Scholar
Hagmann, P., et al. (2008). Mapping the structural core of human cerebral cortex. PLoS Biol. 6(7): e159.Google Scholar
Hall, J. C., Rosbash, M. (1988). Mutations and molecules influencing biological rhythms. Ann Rev Neurosci. 11: 373–93.Google Scholar
Han, W., et al. (2017a). Integrated control of predatory hunting by the central nucleus of the amygdala. Cell. 168(1–2): 311–24.Google Scholar
Han, X., et al. (2017b). Role of dopamine projections from ventral tegmental area to nucleus accumbens and medial prefrontal cortex in reinforcement behaviors assessed using optogenetic manipulation. Metab Brain Dis. doi: 10.1007/s11011-017-0023-3.Google Scholar
Hansen, S., et al. (2017). Testosterone influences volitional, but not reflexive orienting of attention in human males. Physiol Behav. 175: 82–7.Google Scholar
Harris, G. C., Aston-Jones, G. (2006). Arousal and reward: a dichotomy in orexin function. Trends Neurosci. 29(10): 571–7.Google Scholar
Harris, C. W., Edwards, J. L., Baruch, A., et al. (2000). Effects of mental stress on brachial artery flow-mediated vasodilation in healthy normal individuals. Am Heart J. 139: 405–11.Google Scholar
Harrison, L. A., Hurlemann, R., Adolphs, R. (2015). An enhanced default approach bias following amygdala lesions in humans. Psychonomic Sci. 26: 1543–55.Google Scholar
Hasenstaub, A., Shu, Y., Haider, B., et al. (2005). Inhibitory postsynaptic potentials carry synchronized frequency information in active cortical networks. Neuron. 47: 423–35.Google Scholar
Haubensak, W., et al. (2010). Genetic dissection of an amygdala microcircuit that gates conditioned fear. Nature. 468: 270–5.Google Scholar
He, B. J. (2014). Scale-free brain activity: past, present, and future. Trends Cogn Sci. 18(9): 480–7.Google Scholar
He, B. J., (2008). Electrophysiological correlates of the brain's intrinsic large-scale functional architecture. PNAS. 105: 16039–44.Google Scholar
Hebb, D. O. (1955). Drives and the CNS. Psych Rev. 62: 243–54.Google Scholar
Heesink, L., et al. (2017). Anger and aggression problems in veterans are associated with an increased acoustic startle reflex. Biol Psychol. 123: 119–25.Google Scholar
Hemmings, H. C., Hopkins, P. M. (2006). Foundations of Anesthesia (2nd edition). Philadelphia, PA: Mosby/Elsevier.Google Scholar
Herbert, H., Saper, C. B. (1992). Organization of medullary adrenergic and noradrenergic projections to the periaqueductal gray matter in the rat. J Comp Neurol. 315(1): 3452.Google Scholar
Hermann, G. E., Rogers, R. C. (1985). Convergence of vagal and gustatory afferent input within the parabrachial nucleus of the rat. J Auton Nerv Syst. 13(1): 117.Google Scholar
Herold, K. F., Andersen, O. S., Hemmings, H. C. Jr. (2017). Divergent effects of anesthetics on lipid bilayer properties and sodium channel function. Eur Biophys J. doi: 10.1007/s00249-017-1239-1.CrossRefGoogle ScholarPubMed
Hesse, J., Gross, T. (2014). Self-organized criticality as a fundamental property of neural systems. Front Syst Neurosci. 8: 114.Google Scholar
Heyne, H. O., Lautenschläger, S., Nelson, R., et al. (2014) Genetic influences on brain gene expression in rats selected for tameness and aggression. Genetics. 198: 12771290Google Scholar
Hill, C. (2009). Consciousness. Cambridge: University of Cambridge Press.Google Scholar
Hiroi, R., Neumaier, J. F. (2006). Differential effects of ovarian steroids on anxiety versus fear as measured by open field test and fear-potentiated startle. Behav Brain Res. 166(1): 93100.Google Scholar
Hobson, J. A., Scheibel, A. B. (1980). The brainstem core: sensorimotor integration and behavioral state control. Neurosci Res Program Bull. 18(1): 1173.Google Scholar
Hodgins, M. B., Spike, R. C., Mackie, R. M., MacLean, A. B. (1998). An immunohistochemical study of androgen, oestrogen and progesterone receptors in the vulva and vagina. British J Obstet Gynaecol. 105: 216–22.Google Scholar
Holder, M. K., et al. (2010). Methamphetamine facilitates female sexual behavior and enhances neuronal activation in the medial amygdala and ventromedial nucleus of the hypothalamus. Psychoneuroendocrinology. 35(2): 197208.Google Scholar
Holder, M. K., Mong, J. A. (2010). Methamphetamine enhances paced mating behaviors and neuroplasticity in the medial amygdala of female rats. Horm Behav. 58(3): 519–25.CrossRefGoogle ScholarPubMed
Holder, M. K., Veichweg, S. S., Mong, J. A. (2015). Methamphetamine-enhanced female sexual motivation is dependent on dopamine and progesterone signaling in the medial amygdala. Horm Behav. 67: 111.Google Scholar
Horney, C. J., et al. (2009). Predicting human resting-state functional connectivity from structural connectivity. PNAS. 106: 2035–40.Google Scholar
Horvitz, J. C. (2000). Mesolimbocortical and nigrostriatal dopamine responses to salient non-reward events. Neuroscience. 96(4): 651–6.Google Scholar
Hucho, T. B., Dina, O. A., Kuhn, J., Levine, J. D. (2006). Estrogen controls PKCepsilon-dependent mechanical hyperalgesia through direct action on nociceptive neurons. Eur J Neurosci. 24(2): 527–34.Google Scholar
Hudson, A. E., Calderon, D. P., Pfaff, D. W., Proekt, A. (2014). Recovery of consciousness is mediated by a network of discrete metastable activity states. Proc Natl Acad Sci U S A. 111(25): 9283–8.Google Scholar
Hull, E. (2016). Male sexual behavior. In: Pfaff, D., Joels, M. (Eds.), Hormones, Brain and Behavior. Cambridge: Academic Press/Elsevier, pp. 145.Google Scholar
Hull, E. M., Dominguez, J. M. (2006). Getting his act together: roles of glutamate, nitric oxide, and dopamine in the medial preoptic area Brain Res. 1126(1): 6675.CrossRefGoogle ScholarPubMed
Humphrey, N. (1992). A History of the Mind. New York, NY: Copernicus/Springer Verlag.Google Scholar
Humphrey, N. (2011). Soul Dust: The Magic of Consciousness. Princeton, NJ: Princeton University Press.Google Scholar
Hungs, M., et al. (2001). Identification and functional analysis of mutations in the hypocretin (orexin) genes of narcoleptic canines. Genome Res. 11: 531–9.Google Scholar
Hungs, M., Lin, L., Okun, M., Mignot, E. (2001). Polymorphisms in the vicinity of the hypocretin/orexin are not associated with human narcolepsy. Neurology. 57(10): 1893–5.Google Scholar
Iams, S. G., Wexler, B. C. (1979). Inhibition of the development of spontaneous hypertension in SH rats by gonadectomy or estradiol. J Lab Clin Med. 94(4): 608–16.Google Scholar
Ishizuka, T., Murotani, T., Yamatodani, A. (2012). Action of modafinil through histaminergic and orexinergic neurons. Vitamins Hormones. 89: 259–78.Google Scholar
Ito, K., Yanagihara, M., Imon, H., Dauphin, L., McCarley, R. W. (2002). Intracellular recordings of pontine medial gigantocellular tegmental field neurons in the naturally sleeping cat: behavioral state-related activity and soma size difference in order of recruitment. Neuroscience. 114(1): 2337.Google Scholar
Jack, A. I., et al. (2013). fMRI reveals reciprocal inhibition between social and physical cognitive domains. Neuroimage. 66: 385401.Google Scholar
Jacobs, B. L., Fornal, C. A., (2010). Activity of brain serotonergic neurons in relation to physiology and behavior. In: Muller, C., Jacobs, B. (Eds.), Handbook of Behavioral Neurobiology of Serotonin. San Diego, CA: Elsevier/Academic Press, pp. 153–62.Google Scholar
James, W. The Principles of Psychology (volume 2). New York, NY: Dover (1890, reprinted 1950).Google Scholar
Jankowski, M., Rachelska, G., Donghao, W., et al. (2001). Estrogen receptors activate atrial natriuretic peptide in the rat heart. Proc Natl Acad Sci U S A. 98: 11765–70.Google Scholar
Jasper, H. (Ed.). (1958). Reticular Formation of the Brain. Detroit, MI: Symposium of the Henry Ford Hospital.Google Scholar
Jeong, H. H., et al. (2000). The large scale organization of metabolic networks. Nature. 407: 651–4.Google Scholar
Jing, J., Gillette, R., Weiss, K. R. (2009). Evolving concepts of arousal: insights from simple model systems. Rev Neurosci. 20(5–6): 405–27.Google Scholar
Joel, D., Weiner, I. (1994). The organization of the basal ganglia-thalamocortical circuits: open interconnected rather than closed segregated. Neuroscience. 63(2): 363–79.Google Scholar
Johnson, L. R., Hou, M., Prager, E. M., Ledoux, J. E. (2011). Regulation of the fear network by mediators of stress: norepinephrine alters the balance between cortical and subcortical afferent excitation of the lateral amygdala. Front Behav Neurosci. 23(5): 23. doi: 10.3389/fnbeh.2011.00023.Google Scholar
Jones, B. E. (1987). Retrograde labeling of neurones in the brain stem following injections of [3H]choline into the forebrain of the rat. Exp Brain Res. 65(2): 437–48.Google Scholar
Jones, B. E. (2003). Arousal systems. Front Biosci. 8: s438–51.CrossRefGoogle ScholarPubMed
Jones, B. E. (2005). From waking to sleeping: neuronal and chemical substrates. Trends Pharmacol Sci. 26(11): 578–86.Google Scholar
Jones, B. E., Beaudet, A. (1987) Article I. Retrograde labeling of neurones in the brain stem following injections of [3H]choline into the forebrain of the rat. Exp Brain Res. 65(2): 437–48.Google Scholar
Jones, B. E., Cuello, A. C. (1989). Afferents to the basal forebrain cholinergic cell area from pontomesencephalic – catecholamine, serotonin, and acetylcholine – neurons. Neuroscience. 31: 3761.Google Scholar
Jones, B. E., Yang, T. Z. (1985). The efferent projections from the reticular formation and the locus coeruleus studied by anterograde and retrograde axonal transport in the rat. J Comp Neurol. 242: 5692.Google Scholar
Kagan, J., Snidman, N. (1999). Early childhood predictors of adult anxiety disorders. Biol Psychiatry. 46(11): 1536–41.Google Scholar
Kahneman, D. and Tversky, A. (2011). Thinking Fast and Slow. New York, NY: Farrar, Straus & Giroux.Google Scholar
Kaiser, M. (2011). A tutorial in connectome analysis: topological and spatial features of brain networks. Neuroimage. 57(3): 892907.Google Scholar
Kampfl, A., et al. (1998). The persistent vegetative state after closed head injury: clinical and magnetic resonance imaging findings in 42 patients. J Neurosurg. 88: 809–16.Google Scholar
Kandel, E. (2000). Principles of Neural Science (4th edition). New York, NY: McGraw-Hill.Google Scholar
Kandel, E. (2012). The Age of Insight. NewYork, NY: Random House.Google Scholar
Kapas, L., Obal, F. Jr, Book, A. A., et al. (1996). The effects of immunolesions of nerve growth factor-receptive neurons by 192 IgG-saporin on sleep. Brain Res. 712: 53–9.Google Scholar
Karatsoreos, I., Silver, R. (2007). The neuroendocrinology of the suprachiasmatic nucleus as a conductor of body timer in mammals. Endocrinology. 148: 5640–7.Google Scholar
Karatsoreos, I. N., et al. (2011). Androgens modulate structure and function of the suprachiasmatic nucleus brain clock. Endocrinology. 152(5): 1970–8.Google Scholar
Kaur, S., Junek, A., Black, M. A., Semba, K. (2008). Effects of ibotenate and 192IgG-saporin lesions of the nucleus basalis magnocellularis/substantia innominata on spontaneous sleep and wake states and on recovery sleep after sleep deprivation in rats. J Neurosci. 28(2): 491504.Google Scholar
Kaur, S., et al. (2013). Glutamatergic signaling from the parabrachial nucleus plays a critical role in hypercapnic arousal. J Neurosci. 33(18): 7627–40.Google Scholar
Keenan, D. M., Quinkert, A. W., Pfaff, D. W. (2015). Stochastic modeling of mouse motor activity under deep brain stimulation: the extraction of arousal information. PLoS Comput Biol. 11(2): e1003883. doi: 10.1371/journal.pcbi.1003883.Google Scholar
Keizer, K., Kuypers, H. G. (1989). Distribution of corticospinal neurons with collaterals to lower brain stem reticular formation in cat. Experimental Brain Res. 54(1): 107–20.Google Scholar
Kennedy, A., et al. (2014). Internal states and behavioral decision-making: toward an integration of emotion and cognition. Cold Spring Harb Symp Quant Biol. 79: 199210.Google Scholar
Kidd, P. B., Young, M. W., Siggia, E. D. (2015). Temperature compensation and temperature sensation in the circadian clock. Proc Natl Acad Sci U S A. 112(46): E6284–92.Google Scholar
Kim, J. W., Closs, E. I., Albritton, L. M., Cunningham, J. M. (1991). Transport of cationic amino acids by the mouse ecotropic retrovirus receptor. Nature. 352: 725–8.Google Scholar
King, G. W. (1980). Topology of ascending brainstem projections to nucleus parabrachialis in the cat. J Comp Neurol. 191(4): 615–38.CrossRefGoogle ScholarPubMed
Kitamura, T., et al. (2017). Engrams and circuits crucial for systems consolidation of a memory. Science. 356: 7378.Google Scholar
Koch, C. (2008). In: Squire L. R., et al. (Eds.), Fundamental Neuroscience. San Diego, CA: Academic Press (Elsevier), pp. 1223–35.Google Scholar
Koch, K. (2017). How to make a consciousness meter. Scientific American, November, pp. 2834.Google Scholar
Kojima, T., et al. (2009). Default mode of brain activity demonstrated by positron emission tomography imaging in awake monkeys: higher rest-related than working memory-related activity in medial cortical areas. J Neurosci. 29(46): 14463–71.Google Scholar
Kolber Benedict, J., et al. (2008). Central amygdala glucocorticoid receptor action promotes fear-associated CRH activation and conditioning. Proc Natl Acad Sci U S A. 105(33): 12004–9.Google Scholar
Konopka, R. J., Benzer, S. (1971a). Clock mutants of Drosophila melanogaster. Proc Natl Acad Sci U S A. 68(9): 2112–6.Google Scholar
Konopka, R. J., Benzer, S. (1971b). Clock mutants of Drosophila melanogaster. Proc Natl Acad Sci. 68: 2112–18.Google Scholar
Korn, H., Faber, D. S. (2005). The Mauthner cell half a century later. Neuron. 47: 1328.Google Scholar
Kow, L.-M., Pfaff, D. W. (1973). Effects of estrogen treatment on the size of receptive field and response threshold of pudendal nerve in the female rat. Neuroendocrinology, 13: 299313.Google Scholar
Koyama, M., et al. (2011). Mapping a sensory-motor network onto a structural and functional ground plan in the hindbrain. Proc Natl Acad Sciences U S A. 108(3): 1170–5.Google Scholar
Kreibich, A., et al. (2008). Presynaptic inhibition of diverse afferents to the locus ceruleus by kappa-opiate receptors: a novel mechanism for regulating the central norepinephrine system. J Neurosci. 28(25): 6516–25.Google Scholar
Krout, K. E., Loewy, A. D. (2000). Parabrachial nucleus projections to midline and intralaminar thalamic nuclei of the rat. J Comp Neurol. 428: 475–94.3.0.CO;2-9>CrossRefGoogle ScholarPubMed
Kume, K., Kume, S., Park, S. K., Hirsh, J., Jackson, F. R. (2005). Dopamine is a regulator of arousal in the fruit fly. J Neurosci. 25(32): 7377–84.Google Scholar
Kumral, E., Evyapan, D., Balkõr, K., Kutluhan, S. (2001). Bilateral thalamic infarction. Clinical, etiological and MRI correlates. Acta Neurol Scand. 103: 3542.Google Scholar
Kunkhyen, T., et al. (2017). Optogenetic activation of accessory olfactory bulb input to the forebrain differentially modulates investigation of opposite versus same-sex urinary chemosignals and stimulates mating in male mice. eNeuro. 4(2) e0010–17.Google Scholar
Kuo, H. J., Maslen, C. L., Keene, D. R., Glanville, R. W. (1997). Type VI collagen anchors endothelial basement membranes by interacting with type IV collagen. J Biol Chem. 272: 26522–9.Google Scholar
Kurata, J., Hemmings, H. C. (2015). Memory and awareness in anaesthesia; the 9th international conference. Br J Anesthesiol. 115: Editorial, S1.Google Scholar
LaCroix-Fralish, M. L., Tawfik, V. L., DeLeo, J. A. (2005). The organizational and activational effects of sex hormones on tactile and thermal hypersensitivity following lumbar nerve root injury in male and female rats. Pain. 114(1–2): 7180.Google Scholar
LaLumiere, R. T., McGaugh, J. L., McIntyre, C. K. (2017). Emotional modulation of learning and memory: pharmacological implications. Pharmacological Rev. 69(3): 236–55.Google Scholar
Langton, C. G. (1990). Computation at the edge of chaos. Phys D. 42: 1237.Google Scholar
Lanuza, E., Moncho-Bogani, J., Ledoux, J. E. (2008). Unconditioned stimulus pathways to the amygdala: effects of lesions of the posterior intralaminar thalamus on foot-shock-induced c-Fos expression in the subdivisions of the lateral amygdala. Neuroscience. 155(3): 959–68.Google Scholar
Laufs, H. (2012). Functional imaging of seizures and epilepsy: evolution from zones to networks. Curr Opin Neurol. 25: 194200.Google Scholar
Laumann, T. O., et al. (2017). On the stability of BOLD fMRI correlations. Cerebr Cortex. 27: 4719–32.Google Scholar
Laureys, S. (2016a). Traumatic brain damage: Severe brain damage, coma and disorders of consciousness. In: Pfaff, D. W., Volkow, N. D. (Eds.), Neuroscience in the 21st Century (2nd edition, volume 5). New York, NY: Springer Verlag, pp. 3341–70.Google Scholar
Laureys, S. (2016b). Traumatic Brain Damage: Severe brain damage, coma and disorders of consciousness. In: Pfaff, D., Volkow, N., (Eds.), Neuroscience in the 21st Century (2nd edition, volume 5). New York, NY: Springer, pp. 3341–75.Google Scholar
Laureys, S. (2016). In: Pfaff, D., Volkow, N. (Eds.), Neuroscience in the 21st Century (volume 5), pp. 3341–71.Google Scholar
Laureys, Steven, Schiff., Nicholas D. (2012). Coma and consciousness: paradigms (re)framed by neuroimaging. Neuroimage. 61: 478–91.Google Scholar
Ledo, A., Frade, J., Barbosa, R. M., Laranjinha, J. (2004). Nitric oxide in brain: diffusion, targets and concentration dynamics in hippocampal subregions. Mol Aspects Med. 25: 7589.Google Scholar
LeDoux, J. E. (2000). Emotion circuits in the brain. Ann Rev Physiol. 23: 155–84.Google Scholar
LeDoux, J. E. (2014). Coming to terms with fear. Proc Natl Acad Sci. 111: 2871–8.Google Scholar
LeDoux, J. E., Brown, R. (2017). A higher-order theory of emotional consciousness. PNAS. 114: 2016–25.Google Scholar
Lee, M. G., Manns, I. D., Alonso, A., Jones, B. E. (2004). Sleep–wake related discharge properties of basal forebrain neurons recorded with micropipettes in head-fixed rats. J Neurophysiol. 92: 1182–9.Google Scholar
Lee, M. G., Hassani, O. K., Alonso, A., Jones, B. E. (2005). Cholinergic basal forebrain neurons burst with theta during waking and paradoxical sleep. J Neurosci. 25(17): 4365–9.Google Scholar
Lee, H.S., Kim, M.A., and Waterhouse, B.D., (2005). Retrograde double-labeling study of common afferent projections to the dorsal raphe and the nuclear core of the locus coeruleus in the rat. J Comp Neurol. 481: 179–93.Google Scholar
Lee, A. W., et al. (2008). Estradiol modulation of phenylephrine-induced excitatory responses in ventromedial hypothalamic neurons of female rats. Proc Natl Acad Sci U S A. 105(20): 7333–8.Google Scholar
Lee, H., et al. (2014). Scalable control of mounting and attack by Esr11 neurons in the ventromedial hypothalamus. Nature. 509: 627–32.CrossRefGoogle Scholar
Lenaz, G., Fato, R., Genova, M. L., et al. (2006). Mitochondrial Complex I: structural and functional aspects. Biochim Biophys Acta. 1757: 1406–20.Google Scholar
Leontovich, T. A., Zhukova, G. P. (1963). The specificity of the neuronal structure and topography of the reticular formation in the brain and spinal cord of carnivora. J Comp Neurol. 121: 347–79.Google Scholar
Leresche, N., Lambert, R. C. (2017). GABA receptors and T-type Ca2+ channels crosstalk in thalamic networks. Neuropharmacology. pii:S0028-3908(17)30276-9.Google Scholar
LeSauter, J., et al. (2009). Stomach ghrelin-secreting cells as food-entrainable circadian clocks. Proc Natl Acad Sci U S A. 106(32): 13582–7.Google Scholar
Levenson, R. W. (2003). Blood, sweat and fears: the autonomic architecture of emotion. Ann N Y Acad Sci. 1000: 348–66.Google Scholar
Li, H., Satinoff, E. (1996). Body temperature and sleep in intact and ovariectomized female rats. Am J Physiol. 271(6 Pt 2): R1753–8.Google Scholar
Li, S. B., Giardino, W. J., de Lecea, L. (2017). Hypocretins and arousal. Curr Top Behav Neurosci. 33: 93104.Google Scholar
Li, S. B., Jones, J. R., de Lecea, L. (2016). Hypocretins, neural systems, physiology, and psychiatric disorders. Curr Psychiatry Rep. 18(1): 714. doi: 10.1007/s11920-015-0639-0Google Scholar
Li, A., et al. (2017). The fundamental advantages of temporal networks. Science. 358: 1042–52.Google Scholar
Liao, Y., Smyth, G. K., Shi, W. (2014). featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics. 30: 923–30.Google Scholar
Lin, D., Boyle, M. P., Dollar, P., et al. (2011). Functional identification of an aggression locus in the mouse hypothalamus. Nature. 470(7333): 221–6.Google Scholar
Lin, L., Faraco, J., Li, R., et al. (1999). The sleep disorder canine narcolepsy is caused by a mutation in the hypocretin (orexin) receptor 2 gene. Cell. 98(3): 365–76.CrossRefGoogle ScholarPubMed
Lin, J. S., Anaclet, C., Sergeeva, O. A., and Haas, H. L. (2011). The waking brain: an update. Cell Mol Life Sci. 68(15): 2499–512.Google Scholar
Lindsley, D. B., Bowden, J., Magoun, H. W. (1949). Effect upon the EEG of acute injury to the brain stem activating system. Electroencephalogr Clin Neurophysiol. 1: 475–86.Google Scholar
Liu, X., et al. (2016). Development of electrophysiological properties of nucleus gigantocellularis neurons correlated with increased CNS arousal. Dev Neurosci. 38(4): 295310.Google Scholar
Liu, Z.-P., et al. (2017). Delta subunit containing GABA-A receptor disinhibits lateral amygdala and facilitates fear expression in mice. Biol Psychiatry. 81: 9901002.Google Scholar
Liu, Y. Y., Barabasi, A.-L. Network science. Rev Mod Phys. 88(3): 035006–64.Google Scholar
Llinas, R. R., Steriade, M. (2006). Bursting of thalamic neurons and states of vigilance. J Neurophysiol. 95: 3297–308.Google Scholar
Loh, S. Y., Salleh, N. (2017). Influence of testosterone on mean arterial pressure: a physiological study in male and female normotensive WKY and hypertensive SHR rats. Physiol Int. 104(1): 2534.Google Scholar
Loughlin, S. E., Fallon, J. H. (1982). Mesostriatal projections from ventral tegmentum and dorsal raphe: cells project ipsilaterally or contralaterally but not bilaterally. Neurosci Lett. 32(1): 11–6.Google Scholar
Lövblad, K. O., Bassetti, C., Mathis, J., Schroth, G. (1997). MRI of paramedian thalamic stroke with sleep disturbance. Neuroradiology. 39(10): 693–8.Google Scholar
Love, M. I., Huber, W., Anders, S. (2014). Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15: 550.Google Scholar
Lovick, T. A. (1993). The periaqueductal gray-rostral medulla connection in the defence reaction: efferent pathways and descending control mechanisms. Behav Brain Res. 58: 1925.Google Scholar
Lovick, T. A. (2016). Central control of visceral pain and urinary tract function. Autonomic Neurosci. 200: 3542.Google Scholar
Lu, J., Greco, M. A., Shiromani, P., Saper, C. B. (2000). Effect of lesions of the ventrolateral preoptic nucleus on NREM and REM sleep. J Neurosci. 20: 3830–42.Google Scholar
Lu, J., Bjorkum, A. A., Xu, M., et al. (2002). Selective activation of the extended ventrolateral preoptic nucleus during rapid eye movement sleep. J Neurosci. 22: 4568–76.Google Scholar
Lu, J., Jhou, T. C., Saper, C. B. (2006). Identification of wake-active dopaminergic neurons in the ventral periaqueductal gray matter. J Neurosci. 26(1): 193202.Google Scholar
Lu, J., Nelson, L. E., Franks, N., et al. (2008). Role of endogenous sleep-wake and analgesic systems in anesthesia. J Comp Neurol. 508: 648–62.Google Scholar
Lu, J., Sherman, D., Devor, M., Saper, C. B. (2006). A putative flip-flop switch for control of REM sleep. Nature. 441: 589–94.Google Scholar
Lu, Z. M., Li, X. F. (2016). Attack vulnerability of network controllability. PLoS One. 11(9): e0162289. doi: 10.1371/journal.pone.0162289.Google Scholar
Luigetti, M., Di Lazzaro, V., Broccolini, A., et al. (2011). Bilateral thalamic stroke transiently reduces arousals and NREM sleep instability. J Neurol Sci. 300: 151–4.Google Scholar
Lund, T. D., Rovis, T., Chung, W. C., Handa, R. J. (2005). Novel actions of estrogen receptor-beta on anxiety-related behaviors. Endocrinology. 146(2): 797807.Google Scholar
Lunga, P., Herbert, J. (2004). 17Beta-oestradiol modulates glucocorticoid, neural and behavioural adaptations to repeated restraint stress in female rats. J Neuroendocrinol. 16(9): 776–85.Google Scholar
Luo, A. H., Aston-Jones, G. (2009). Circuit projection from suprachiasmatic nucleus to ventral tegmental area: a novel circadian output pathway. Eur J Neurosci. 10: 113.Google Scholar
Lutkenhoff, E. S., et al. (2015). Thalamic and extra-thalamic mechanisms of consciousness after severe brain injury. Ann Neurol. 78: 6876.Google Scholar
Ma, S., et al. (2017). Dual-transmitter systems regulating arousal, attention, learning and memory. Neurosci Biobehav Rev. pii:S0149-7634(17)30066-0. doi: 10.1016/j.neubiorev.2017.07.009.Google Scholar
MacLeod, N. K., Mayer, M. L. (1980). Electrophysiological analysis of pathways connecting the medial preoptic area with the mesencephalic central grey matter in rats. J Physiol. 298: 5370.Google Scholar
Magnasco, M. (2003). A wave traveling over a Hopf instability shapes the cochlear tuning curve. Phys Rev Lett. 84: 243–6.Google Scholar
Magnasco, M. O., Piro, O., Cecchi, G. A. (2009). Self-tuned critical anti-Hebbian networks. Phys Rev Lett. 102(25): 258102.Google Scholar
Magoun, H. W. (1958). The Waking Brain (2nd edition). Springfield, IL: Charles C Thomas.Google Scholar
Mann, K., et al. (2016). Immunity around the clock. Science. 354: 9991009.Google Scholar
Manford, M., Andermann, F. (1998). Complex visual hallucinations. Clinical and neurobiological insights. Brain. :1819–40.Google Scholar
Marcus, E. Raichle. (2006). The brain's dark energy. Science. 314: 1249–50.Google Scholar
Marcus, J. N., et al. (2001). Differential expression of orexin receptors 1 and 2 in the rat brain. J Comp Neurol. 435(1): 625.Google Scholar
Marcus, J. N., Aschkenasi, C. J., Lee, C. E., et al. (2001). Differential expression of orexin receptors 1 and 2 in the rat brain. J Comp Neurol. 435(1): 625.Google Scholar
Marin, M. F., et al. (2017). Skin conductance responses and neural activations during fear conditioning and extinction recall across anxiety disorders. JAMA Psychiatry. 74(6): 622–31.Google Scholar
Markowitsch, H. J., Irle, E. (1981). Widespread cortical projections of the ventral tegmental area and of other brain stem structures in the cat. Exp Brain Res. 41: 233–46.Google Scholar
Marshall, W., Albantakis, L., Tononi, G. (2018) Article I. Black-boxing and cause-effect power. PLoS Comput Biol. 14(4): e1006114. doi: 10.1371/journal.pcbi.1006114. eCollection 2018 Apr.Google Scholar
Martin, E. M., Pavlides, C., Pfaff, D. (2010). Multimodal sensory responses of nucleus reticularis gigantocellularis and the responses’ relation to cortical and motor activation. J Neurophysiol. 103(5): 2326–38.Google Scholar
Martin, E. M., Devidze, N., Shelley, D. N., et al. (2011). Molecular and neuroanatomical characterization of single neurons in the mouse medullary gigantocellular reticular neurons. J Comp Neurol. 519(13): 2574–93.Google Scholar
Martin-Alguacil, N., Schober, J. M., Kow, L. M., Pfaff, D. (2006). Arousing properties of the vulvar epithelium. J Urol. 176(2): 456–62.Google Scholar
Martin-Alguacil, N., et al. (2008a). Clitoral sexual arousal: neuronal tracing study from the clitoris through the spinal tracts. J Urol. 180(4): 1241–8.Google Scholar
Martin-Alguacil, N., (2008b). Oestrogen receptor expression and neuronal nitric oxide synthase in the clitoris and prepucial gland structures of mice. BJU Int. 102(11): 1719–23.CrossRefGoogle Scholar
Martin-Alguacil, N., Pfaff, D. W., Kow, L. M., Schober, J. M. (2008c). Oestrogen receptors and their relation to neural receptive tissue of the labia minora. BJU Int. 101(11): 1401–6.Google Scholar
Martin-Alguacil, N., Pfaff, D. W., Shelley, D. N., Schober, J. M. (2008d). Clitoral sexual arousal: an immunocytochemical and innervation study of the clitoris. BJU Int. 101(11): 1407–13.Google Scholar
Mason, P., Leung, C. (1996). Physiological functions of ontomedullary raphe and medial reticular neurons. In: Holstege, G., et al. (Eds.), Progress in Brain Research (volume 107), pp. 269–81.Google Scholar
Mattson, M. P. (2003). Neurobiology of Aggression. Totowa, NJ: Humana Press.Google Scholar
Matuszewich, L., Lorrain, D. S., Hull, E. M. (2000). Dopamine release in the medial preoptic area of female rats in response to hormonal manipulation and sexual activity. Behav Neurosci. 114(4): 772–82.Google Scholar
McBride, R. L., Sutin, J. (1976). Projections of the locus coeruleus and adjacent pontine tegmentum in the cat. J Comp Neurol. 165(3): 265–84.Google Scholar
McCann, J., Miyamoto, S., Boyle, C., Rogers, K. (2007). Healing of hymenal injuries in prepubertal and adolescent girls: a descriptive study. Pediatrics. 119(5): e1094–106.Google Scholar
McCarthy, E. A., et al. (2017). A comparison of the effects of male pheromone priming and optogenetic inhibition of accessory olfactory bulb forebrain inputs on the sexual behavior of estrous female mice. Horm Behav. 89: 104–12.Google Scholar
McCulloch, W. S., Pitts, W. (1943). A logical calculus of the ideas immanent in nervous activity. Bull Math Biophys. 5: 115–33.Google Scholar
McDuffie, J. E., Coaxum, S. D., Maleque, M. A. (1999). 5-Hydroxytryptamine evokes endothelial nitric oxide synthase activation in bovine aortic endothelial cell cultures. Proc Soc Exp Biol Med. 221: 386–90.Google Scholar
Melloni, R. H., Ricci, L. A. (2010). Adolescent exposure to anabolic/androgenic steroids and the neurobiology of offensive aggression. Horm Behav. 58: 177–91.Google Scholar
Mena-Segovia, J., Bolam, J. P. (2017). Rethinking the pedunculopontine nucleus: from cellular organization to function. Neuron. 94(1): 718.Google Scholar
Mendelsohn, M. E., Karas, R. H. (1999). The protective effects of estrogen on the cardiovascular system. New Engl J Med. 340: 1801–11.Google Scholar
Menétrey, D., De Pommery, J. (1991). Origins of spinal ascending pathways that reach central areas involved in visceroception and visceronociception in the rat. Eur J Neurosci. 3(3): 249–59.Google Scholar
Meston, C. M., Moe, I. V., Gorzalka, B. B. (1996). Effects of sympathetic inhibition on receptive, proceptive, and rejection behaviors in the female rat. Physiol Behav. 59: 537–42.Google Scholar
Metzinger, T. (2002). Neural Correlates of Consciousness. Cambridge, MA: MIT Press.Google Scholar
Miczek, K. A., Faccidomo, S., De Almeida, R. M., et al. (2004) Article I. Escalated aggressive behavior: new pharmacotherapeutic approaches and opportunities. Ann NY Acad Sci. 1036: 336–55.Google Scholar
Miller, R. D. (Ed.). (2005). Miller's Anesthesia (6th edition) Philadelphia, PA: Churchill Livingstone/Elsevier.Google Scholar
Minert, A., Yatziv, S.-L., Devor, M. (2017). Location of the mesopontine neurons responsible for maintenance of anesthetic loss of consciousness. J Neurosci. 37(38): 9320–31.Google Scholar
Misonou, H., Mohapatra, D. P., Trimmer, J. S. (2005). Kv2.1: a voltage-gated k+ channel critical to dynamic control of neuronal excitability. Neurotoxicology. 26: 743–52.Google Scholar
Mitchell, C. L. Kaelber, W. W. (1967). Unilateral vs bilateral medial thalamic lesions and reactivity to noxious stimuli. Arch Neurol. 17(6): 653–60.Google Scholar
Mitra, C., et al. (2017). Multiple-node basin stability in complex dynamical networks. Phys Rev E. 95(3–1): 032317. doi: 10.1103.Google Scholar
Mlinar, B., et al. (2016). Firing properties of genetically identified dorsal raphe serotonergic neurons in brain slices. Front Cell Neurosci. 10: 195. Doi: 10.3389/fncel.2016.00195.Google Scholar
Model, Z., et al. (2015). Suprachiasmatic nucleus as the site of androgen action on circadian rhythms. Horm Behav. 73: 17.Google Scholar
Moga, M. M., Herbert, H., Hurley, K. M., et al. (1990). Organization of cortical, basal forebrain, and hypothalamic afferents to the parabrachial nucleus in the rat. J Comp Neurol. 295: 624–61.Google Scholar
Mong, J. A., et al. (2003a). Estradiol differentially regulates lipocalin-type prostaglandin D synthase transcript levels in the rodent brain: evidence from high-density oligonucleotide arrays and in situ hybridization. Proc Natl Acad Sci U S A. 100(1): 318–23.Google Scholar
Mong, J. A., Devidze, N., Goodwillie, A., Pfaff, D. W. (2003b). Reduction of lipocalin-type prostaglandin D synthase in the preoptic area of female mice mimics estradiol effects on arousal and sex behavior. Proc Natl Acad Sci U S A. 100(25): 15206–11.Google Scholar
Mong, J. A., et al. (2011). Sleep, rhythms, and the endocrine brain: influence of sex and gonadal hormones. J Neurosci. 31(45): 16107–16.Google Scholar
Monti, M. M., et al. (2015). Thalamo-frontal connectivity mediates top-down cognitive functions in disorders of consciousness. Neurology. 84: 167–73.Google Scholar
Monti Martin, M. (2016). Non-invasive ultrasonic thalamic stimulation in disorders of consciousness after severe brain injury: a first-in-man report. Brain Stimul. 9: 940–1.Google Scholar
Moore, R. Y., Abrahamson, E. A., Van Den Pol, A. (2001). The hypocretin neuron system: an arousal system in the human brain. Arch Ital Biol. 139(3): 195205.Google Scholar
Morgan, M., Pfaff, D. (2001) Effects of estrogen on activity and fear-related behaviors in mice. Horm. Behav. 40: 472482Google Scholar
Morin, L.P. (2013) Neuroanatomy of the extended circadian rhythm system. Exp Neurol. 243: 420. doi: 10.1016/j.expneurol.2012.06.026. Epub 2012 Jul 2.Google Scholar
Morison, R. S., Dempsey, E. W. (1942). A study of thalamo-cortical relations. Am J Physiol. 135: 281–92.Google Scholar
Morrell, J. I., Pfaff, D. W. (1983). Retrograde hrp identification of neurons in the rhombencephalon and spinal cord of the rat that project to the dorsal mesencephalon. Am J Anat. 167: 229–40.Google Scholar
Morrell, J. I., Greenberger, L. M., Pfaff, D. W. (1981). Hypothalamic, other diencephalic, and telencephalic neurons that project to the dorsal midbrain. J Comp Neurol. 201: 589620.Google Scholar
Moruzzi, G., Magoun, H. (1949). Brain stem reticular formation and activation of the EEG. Electroencephalogr Clin Neurophysiol. 1: 455–73. [PubMed]Google Scholar
Motta, S. C., et al. (2017). The periaqueductal gray and primal emotional processing critical to influence complex defensive responses, fear learning and reward seeking. Neurosci Biobehav Rev. 76(Pt A): 3947.Google Scholar
Motter, A. E. (2004). Cascade control and defense in complex networks. Phys Rev Lett. 93(9): 098701.Google Scholar
Mountcastle, V. B. (1974). Mountcastle, VB Medical Physiology (13th edition). St Louis, MO: Mosby, pp. 254–84.Google Scholar
Mukouyama, Y. S., Gerber, H. P., Ferrara, N., Gu, C., Anderson, D. J. (2005). Peripheral nerve-derived VEGF promotes arterial differentiation via neuropilin 1-mediated positive feedback. Development. 132(5): 941–52.Google Scholar
Munk, M. H., Roelfsema, P. R., Konig, P., Engel, A. K., Singer, W. (1996). Role of reticular activation in the modulation of intracortical synchronization. Science. 272: 271–4.Google Scholar
Muoio, V., Persson, P. B., Sendeski, M. M. (2014). The neurovascular unit – concept review. Acta Physiol (Oxf). 210: 790–8.Google Scholar
Murphy, C. P., et al. (2017). MicroRNA-mediated rescue of fear extinction memory by miR 144-3p in extinction-impaired mice. Biol Psychiatry. 81: 979–89.Google Scholar
Muschamp, J. W., et al. (2007). A role for hypocretin (orexin) in male sexual behavior. J Neuroscience. 27(11): 2837–45.Google Scholar
Myers, B., et al. (2017). Ascending mechanisms of stress integration: implications for brainstem regulation of neuroendocrine and behavioral stress responses. Neurosci Biobehav Rev. 74(Pt B): 366–75.Google Scholar
Nakajima, M., Halassa, M. M. (2017). Thalamic control of functional cortical connectivity. Curr Opin Neurobiol. 44: 127–31.Google Scholar
Nauta, W. J. H. (1946). Hypothalamic regulation of sleep in rats. Experimental study. J Neurophysiol. 9: 285316.Google Scholar
Nauta, W. J. H., Kuypers, H. G. J. M. (1958). Some ascending pathways in the brainstem reticular formation. In: Jasper, H. (Ed.), The Reticular Formation of the Brain. Boston, MA: Little Brown, pp. 330.Google Scholar
Nautiyal, K. M., Hen, R. (2017). Serotonin receptors in depression: from A to B. F1000Res. 6: 123. doi: 10.12688/f1000research.9736.1.Google Scholar
Nectow, A. R., Ekstrand, M. I., Friedman, J. M. (2015). Molecular characterization of neuronal cell types based on patterns of projection with Retro-TRAP. Nat Protoc. 10: 1319–27.Google Scholar
Nelson, R. (Ed.). (2006). Biology of Aggression. Oxford: Oxford University Press.Google Scholar
Ng, R. C., et al. (2010). Pharmacologic treatment for postpartum depression: a systematic review. Pharmacotherapy. 30(9): 928–41.Google Scholar
Nickenig, G., Strehlow, K., Wassmann, S., et al. (2000). Differential effects of estrogen and progesterone on AT(1) receptor gene expression in vascular smooth muscle cells. Circulation. 102: 1828–33.Google Scholar
Nicoll, R. A. (2017). A brief history of long-term potentiation. Neuron. 93: 281–90.Google Scholar
Nishino, S., et al. (2000). Hypocretin (orexin) transmission is defective in human narcolepsy. Lancet. 355: 3940.Google Scholar
Nishino, S., (2001). Low cerebrospinal fluid hypocretin (Orexin) and altered energy homeostasis in human narcolepsy. Ann Neurol. 50(3): 381–8.Google Scholar
Nomura, M., Durback, L., Chan, J., et al. (2002). Genotype/age interactions on aggressive behavior in gonadally intact estrogen receptor β knockout (βERKO) male mice. Horm Behav. 41(3): 288–96.Google Scholar
Norton, L., et al. (2012). Disruptions of functional connectivity in the default mode network of comatose patients. Neurology. 78: 175–81.Google Scholar
Nyberga, L., et al. (2010). Consciousness of subjective time in the brain. PNAS. 107: 22356–9.Google Scholar
Ogawa, S., et al. (1998a). Roles of estrogen receptor-alpha gene expression in reproduction-related behaviors in female mice. Endocrinology. 139(12): 5070–81.Google Scholar
Ogawa, S., (1998b). Modifications of testosterone-dependent behaviors by estrogen receptor-alpha gene disruption in male mice. Endocrinology. 139: 5058–69.Google Scholar
Ogawa, S., Chan, J., Gustafsson, J. A., Korach, K. S., Pfaff, D. W. (2003). Estrogen increases locomotor activity in mice through estrogen receptor alpha: specificity for the type of activity. Endocrinology. 144(1): 230–9.Google Scholar
Ogawa, S., Choleris, E., Pfaff, D. (2004). Genetic influences on aggressive behaviors and arousability in animals. Ann N Y Acad Sci. 1036: 257–66.Google Scholar
Ogawa, S., Nomura, M., Choleris, E., Pfaff, D., in Nelson (op cit., 2006). The roles of estrogen receptors in the regulation of aggressive behaviors. pp. 231250Google Scholar
Oloyo, A. K., et al. (2016). Orchidectomy attenuates high-salt diet-induced increases in blood pressure, renovascular resistance, and hind limb vascular dysfunction: role of testosterone. Clin Exp Pharmacol Physiol. 43(9): 825–33.Google Scholar
Ott, E., et al. (1994). Coping with Chaos: Analysis of Chaotic Data and the Exploitation of Chaotic Systems. New York, NY: Wiley.Google Scholar
Owen, A. M., Coleman, M. R. (2008a). Functional neuroimaging of the vegetative state. Nat Rev Neurosci. 9(3): 235–43.Google Scholar
Owen, A. M., Coleman, M. R. (2008b). Using neuroimaging to detect awareness in disorders of consciousness. Funct Neurol. 23(4): 189–94.Google Scholar
Owen, A. M., et al. (2006). Detecting awareness in the vegetative state. Science. 313(5792): 1402.Google Scholar
Pace-Schott, E. F., et al. (2008). In: Squire, L. R., et al. (Eds.), Fundamental Neuroscience. San Diego, CA: Academic Press (Elsevier), pp. 958–85.Google Scholar
Palmer, R. M., Ashton, D. S., Moncada, S. (1988). Vascular endothelial cells synthesize nitric oxide from L-arginine. Nature. 333: 664–6.Google Scholar
Panksepp, J. (1998). Affective Neuroscience. New York, NY: Oxford University Press.Google Scholar
Panula, P., et al. (2015). International union of basic and clinical pharmacology. XCVIII. Histamine receptors. Pharmacol Rev. 67(3): 601–55.Google Scholar
Papka, R. E., Srinivasan, B., Miller, K. E., Hayashi, S. (1997). Localization of estrogen receptor protein and estrogen receptor messenger RNA in peripheral autonomic and sensory neurons. Neuroscience. 79: 1153–63.Google Scholar
Park, K. K., Liu, K., Hu, Y., et al. (2008). Promoting axon regeneration in the adult CNS by modulation of the PTEN/mTOR pathway. Science. 322: 963–6.Google Scholar
Parvizi, J., Damasio, A. R. (2003). Neuroanatomical correlates of brainstem coma. Brain. 126: 1524–36.Google Scholar
Paulauskis, J. D., Sul, H. S. (1989). Structure of mouse fatty acid synthase mRNA. Identification of the two NADPH binding sites. Biochem Biophys Res Commun. 158: 690–5.Google Scholar
Paul, M. J., Indic, P., Schwartz, W. J. (2011) Article I. A role for the habenula in the regulation of locomotor activity cycles. Eur J Neurosci. 34(3): 478-88. doi: 10.1111/j.1460-9568.2011.07762.x. Epub 2011 Jul 21.Google Scholar
Pauls, S. D., et al. (2016). Deconstructing circadian rhythmicity with models and manipulations. Trends Neurosci. 39(6): 405–19.Google Scholar
Pereira de Vasconcelos, A., Cassel, J. C. (2015). The nonspecific thalamus: a place in a wedding bed for making memories last? Neurosci Biobehav Rev. 54: 175–96.Google Scholar
Perkins, E., May, P. J., Warren, S. (2014). Feed-forward and feedback projections of midbrain reticular formation neurons in the cat. Front Neuroanat. 7: 5571.Google Scholar
Peschanski, M., Besson, J. M. (1984). A spino-reticulo-thalamic pathway in the rat: an anatomical study with reference to pain transmission. Neuroscience. 12(1): 165–78.Google Scholar
Pessoa, L., Ungerleider, L. G. (2004). Neuroimaging studies of attention and the processing of emotion-laden stimuli. Prog Brain Res. 144: 171–82.Google Scholar
Peter, Kuppens P., et al. (2016). The relation between valence and arousal in subjective experience varies with personality and culture. J Pers. 10: 1111–25.Google Scholar
Peterfi, L., et al. (2004). Fos-immunoreactivity in the hypothalamus: dependency on the diurnal rhythm, sleep, gender, and estrogen. Neuroscience. 124: 695707.Google Scholar
Peterson, B. W. (1979) Reticulospinal projections to spinal motor nuclei. Annu Rev Physiol. 41: 127–40.Google Scholar
Peterson, B. W., Pitts, N. G., Fukushima, K. (1979). Reticulospinal connections with limb and axial motoneurons. Exp Brain Res. 36(1): 120.Google Scholar
Peyron, C., et al. (2000). A mutation in early onset narcolepsy and a generalized absence of hypocretin peptides in human narcoleptic brains. Nat Med. 6: 991–7.Google Scholar
Pfaff, D. (1999). Drive. Cambridge, MA: MIT Press.Google Scholar
Pfaff, D. (2006). Brain Arousal and Information Theory. Cambridge, MA: Harvard University Press.Google Scholar
Pfaff, D. (2014). Altruistic Brain Theory. New York, NY: Oxford University Press.Google Scholar
Pfaff, D., Young, L. (2014). Frontiers in Neuroendocrinology: Sex Differences in Neurological and Psychiatric Disorders. Amsterdam: Elsevier.Google Scholar
Pfaff, D. W. (2006). Brain Arousal and Information Theory: Neural and Genetic Mechanisms (Cambridge, MA: Harvard University Press).Google Scholar
Pfaff, D. W. (2017). How the Vertebrate Brain Regulates Behavior. Cambridge, MA: Harvard University Press.Google Scholar
Pfaff, D. W., Banavar, J. R. (2007). A theoretical framework for CNS arousal. BioEssays. 29(8): 803–10.Google Scholar
Pfaff., D. W., Kieffer, B. L. (Eds). (2008). Molecular and biophysical mechanisms of arousal, alertness and attention. Ann N Y Acad Sci. 1129: 115.Google Scholar
Pfaff, D. W., Baum, M. (2017). Hormone-dependent medial preoptic/lumbar spinal cord/autonomic coordination supporting male sexual behaviors. Mol Cell Endocrinol. 467: 2130. doi: 10.1016/j.mce.2017.10.018.Google Scholar
Pfaff, D. W., et al. (2004). Principles of Hormone/Behavior Relations (2nd edition, 2018). San Diego, CA: Academic Press/Elsevier.Google Scholar
Pfaff, D. W., Westberg, L., Kow, L. M. (2005). Generalized arousal of mammalian central nervous systems. J Comp Neurol. 493(1): 8691.Google Scholar
Pfaff, D. W., Kieffer, B. L., Swanson, L. W. (2008). Mechanisms for the regulation of state changes in the CNS. An introduction. Ann N Y Acad Sci. 1129: 17.Google Scholar
Pfaff, D. W., Rapin, I., Goldman, S. (2011). Male predominance in autism: neuroendocrine influences on arousal and social anxiety. Autism Res. 4(3): 163–76.Google Scholar
Pfaff, D. W., Martin, E. M., Faber, D. (2012). Origins of arousal: roles for medullary reticular neurons. Trends Neurosci. 35(8): 468–76.Google Scholar
Pfaff, D. W., Gagnidze, K., Hunter, R. G. (2017). Molecular endocrinology of female reproductive behavior. Mol Cell Endocrinol. 467: 1420. doi: 10.1016/j.mce.2017.10.019.Google Scholar
Pfaus, J. G. (2010). Inhibitory and disinhibitory effects of psychomotor stimulants and depressants on the sexual behavior of male and female rats. Horm Behav. 58(1): 163–76.Google Scholar
Pfaus, J. G., Pfaff, D. W. (1992). Mu, delta, and kappa opioid receptor agonists selectively modulate sexual behaviors in the female rat: differential dependence on progesterone. Horm Behav. 26: 457–73.Google Scholar
Posner, J., Russell, J. A., Peterson, B. S. (2005). The circumplex model of affect: an integrative approach to affective neuroscience, cognitive development, and psychopathology. Dev Psychopathol. 17(3): 715–34.Google Scholar
Posner, J. B., Saper, C. B., Schiff, N. D., Plum, F. (2007). Diagnosis of Stupor and Coma (volume 4). New York, NY: Oxford University Press, pp. 2934.Google Scholar
Postfai, M., Barabasi, A.-L. (2017). Properties of scale-free networks. Phys Rev E. 94(3): 032316.Google Scholar
Presta, A., Liu, J., Sessa, W. C., Stuehr, D. J. (1997). Substrate binding and calmodulin binding to endothelial nitric oxide synthase coregulate its enzymatic activity. Nitric Oxide. 1: 7487.Google Scholar
Proekt, A., Banavar, J. R., Maritan, A., Pfaff, D. W. (2012). Scale invariance in the dynamics of spontaneous behavior. Proc Natl Acad Sci. 109(26): 10564–9.Google Scholar
Prouty, E. W., Waterhouse, B. D., Chandler, D. J. (2017). Corticotropin releasing factor dose-dependently modulates excitatory synaptic transmission in the noradrenergic nucleus locus coeruleus. Eur J Neurosci. 45(5): 712–22.Google Scholar
Puzzo, D., Staniszewski, A., Deng, S. X., et al. (2009). Phosphodiesterase 5 inhibition improves synaptic function, memory, and amyloid-beta load in an Alzheimer's disease mouse model. J Neurosci. 29: 8075–86.Google Scholar
Quinkert, A. W., Pfaff, D. W. (2012). Temporal patterns of deep brain stimulation generated with a true random number generator and the logistic equation: effects on CNS arousal in mice. Behav Brain Res. 229(2): 349–58.Google Scholar
Quinkert, A. W., Schiff, N. D., Pfaff, D. W. (2010). Temporal patterning of pulses during deep brain stimulation affects central nervous system arousal. Behav Brain Res. 214(2): 377–85.Google Scholar
Quinkert, A. W., Vimal, V., Reeke, G., et al. (2011). Quantitative descriptions of generalized arousal, an elementary function of the vertebrate brain. Proc Natl Acad Sci. 108(Suppl. 3): 15617–23.Google Scholar
Quirk, G. J. (2016). Fear. In: Pfaff, D. W., Volkow, N. C. (Eds.), Neuroscience in the 21st Century (2nd edition). New York, NY: Springer, pp. 2412–35.Google Scholar
Rahman, N., et al. (2018). Mathematical description of the phase transition from low to high behavioral activity. Nature (submitted).Google Scholar
Raichle, M. E. (2010). Two views of brain function. Trends Cogn Sci. 14: 180–90.Google Scholar
Raichle, M. E., (2001). A default mode of brain function. Proc Natl Acad Sci U S A. 98(2): 676–82.Google Scholar
Raimondo, F., et al. (2017). Brain–heart interactions reveal consciousness in noncommunicating patients. Ann Neurol. 82(4): 578–91.Google Scholar
Ramon-Moliner, E., Nauta, W. J. (1966). The isodendritic core of the brain stem. J Comp Neurol. 126: 311–35.Google Scholar
Ranson, S. W. (1939). Somnolence caused by hypothalamic lesions in monkeys. Arch Neurol Psychiatr. 41: 123.Google Scholar
Rasmussen, J. J., et al. (2017). Increased blood pressure and aortic stiffness among abusers of anabolic androgenic steroids: potential effect of suppressed natriuretic peptides in plasma? J Hypertension. doi: 10.1097/HJH.0000000000001546.Google Scholar
Ravasz, E., Barabási, A. L. (2003). Hierarchical organization in complex networks. Phys Rev E Stat Nonlin Soft Matter Phys. 67(2 Pt 2): 026112.Google Scholar
Ray, S., Maunsell, J. H. (2015). Do gamma oscillations play a role in cerebral cortex? Trends Cogn Sci. 19(2): 7885.Google Scholar
Reddy, P., et al. (1984). Molecular analysis of the period locus in Drosophila melanogaster and identification of a transcript involved in biological rhythms. Cell. 38(3): 701–10.Google Scholar
Reich, P. B., Tjoelker, M. G., Machado, J-L., and Oleksyn, J. (2006). Universal scaling of respiration, metabolism, size and nitrogen in plants. Nature. 439: 457–61.Google Scholar
Reyes, B. A., Valentino, R. J., Xu, G., Van Bockstaele, E. J. (2005). Hypothalamic projections to locus coeruleus neurons in rat brain. Eur J Neurosci. 22(1): 93106.Google Scholar
Rhodes, C. H., Morrell, J. I., Pfaff, D. W. (1982). Estrogen-concentrating neurophysin-containing hypothalamic magnocellular neurons in the vasopressin-deficient (Brattleboro) rat: a study combining steroid autoradiography and immunocytochemistry. J Neurosci. 2: 1718–24.Google Scholar
Ribeiro, A. C., et al. (2007). Two forces for arousal: Pitting hunger versus circadian influences and identifying neurons responsible for changes in behavioral arousal. Proc Natl Acad Sci U S A. 104(50): 20078–83.Google Scholar
Ribeiro, A. C., Pfaff, D. W., Devidze, N. (2009). Estradiol modulates behavioral arousal and induces changes in gene expression profiles in brain regions involved in the control of vigilance. Eur J Neurosci. 29(4): 795801.Google Scholar
Rohaut, B., Naccache, L. (2017). Disentangling conscious from unconscious cognitive processing with event-related EEG potentials. Rev Neurologie (Paris). 173(7–8): 521–8.Google Scholar
Romanov, R. A., et al. (2017). Molecular interrogation of hypothalamic organization reveals distinct dopamine neuronal subtypes. Nat Neurosci. 20(2): 176–88.Google Scholar
Roozendaal, B., et al. (1999). Basolateral amygdala noradrenergic influence enables enhancement of memory consolidation induced by hippocampal glucocorticoid receptor activation. Proc Natl Acad Sci U S A. 96(20): 11642–7.Google Scholar
Rowe, D. C., Plomin, R. (1977). Temperament in early childhood. J Pers Assess. 41(2): 150–6.Google Scholar
Russell, J. A. (1980). A circumplex model of affect. J Pers Social Psychol. 39: 1151–78.Google Scholar
Saez, L., Young, M. W. (1988). In situ localization of the per clock protein during development of Drosophila melanogaster. Mol Cell Biol. 8(12): 5378–85.Google Scholar
Sakai, K., et al. (1976). Afferent projections to the locus coeruleus nucleus in the cat. Study by the horseradish peroxidase technic. C R Seances Soc Biol Fil. 170(1): 115–19.Google Scholar
Sakurai, T., et al. (1998). Orexin and orexin receptors: a family of hypothalamic neuropeptides and G-protein coupled receptors that regulate feeding behavior. Cell. 92: 573–85.Google Scholar
Sand, R. M., et al. (2017). Isoflurane modulates activation and inactivation gating of the prokaryotic Na+ channel NaChBac. J Gen Physiol. 149(6): 623–38.Google Scholar
Saper, C., et al. (2010). Sleep state switching. Neuron. 68: 1023–42.Google Scholar
Saper, C. B. (1982). Reciprocal parabrachial–cortical connections in the rat. Brain Res. 242(1): 3340. [PubMed]Google Scholar
Saper, C. B. (1984). Organization of cerebral cortical afferent systems in the rat. II. Magnocellular basal nucleus. J Comp Neurol. 222: 313–42.Google Scholar
Saper, C. B. (1985). Organization of cerebral cortical afferent systems in the rat. II. Hypothalamocortical projections. J Comp Neurol. 237: 2146.Google Scholar
Saper, C. B. (2000). Brainstem modulation of sensation, movement and consciousness. In: Kandel, E. R., et al. (Eds.), Principles of Neural Science (4th edition). New York, NY: McGraw-Hill, pp. 889910.Google Scholar
Saper, C. B., Loewy, A. D. (1980). Efferent connections of the parabrachial nucleus in the rat. Brain Res. 197: 291317.Google Scholar
Saper, C. B., Loewy, A. D. (2016). Commentary on: efferent connections of the parabrachial nucleus in the rat. Brain Res. 1645: 1517.Google Scholar
Saper, C. B., Chou, T. C., Scammell, T. E. (2001). The sleep switch: hypothalamic control of sleep and wakefulness. Trends Neurosci. 24(12): 726–31.Google Scholar
Saper, C. B., Scammell, T. E., Lu, J. (2005a). Hypothalamic regulation of sleep and circadian rhythms. Nature. 437: 1257–63.Google Scholar
Saper, C. B., Scammell, T. E., Lu, J. (2005b). Hypothalamic regulation of sleep and circadian rhythms. Nature. 437: 1257–61.Google Scholar
Saper, C. B., Fuller, P. M., Pedersen, N. P., Lu, J., Scammell, T. E. (2010). Sleep state switching. Neuron. 68(6): 1023–42.Google Scholar
Saper, C. B., Loewy, A. D., Swanson, L. W., Cowan, W. M. (1976) Article I. Direct hypothalamo-autonomic connections. Brain Res. 117(2): 305–12.Google Scholar
Sara, S. J., Bouret, S. (2012). Orienting and reorienting: the locus coeruleus mediates cognition through arousal. Neuron. 76: 130–41.Google Scholar
Sato, T. N., Tozawa, Y., Deutsch, U., et al. (1995). Distinct roles of the receptor tyrosine kinases Tie-1 and Tie-2 in blood vessel formation. Nature. 376: 70–4.Google Scholar
Scammell, T. E., et al. (2000). Hypothalamic arousal regions are activated during modafinil-induced wakefulness. J Neurosci. 20(22): 8620–8.Google Scholar
Schaafsma, S., Pfaff, D. (2014). Etiologies underlying sex differences in Autism Spectrum Disorders. Front Neuroendocrinol. 35: 255–72.Google Scholar
Scheibel, M., Scheibel, A. (1961). Structural substrates for integrative patterns in the brain stem reticular core. In: Jasper, H. (Ed.), Reticular Formation of the Brain. Boston, MA: Little, Brown, pp. 3168.Google Scholar
Scheibel, M. E., Scheibel, A. B. (1961). On circuit patterns of the brain stem reticular core. Annals N Y Acad Sci. 89: 857–65.Google Scholar
Schiff, N. D. (2005). Modeling the minimally conscious state: measurements of brain function and therapeutic possibilities. Prog Brain Res. 150: 473–93.Google Scholar
Schiff, N. D. (2008). Central thalamic contributions to arousal regulation and neurological disorders of consciousness. Ann N Y Acad Sci. 1129: 105–18.Google Scholar
Schiff, N. D. (2009). Recovery of consciousness after brain injury: a mesocircuit hypothesis. Trends Neurosci. 33: 19.Google Scholar
Schiff, N. D. (2010). Recovery of consciousness after brain injury: a mesocircuit hypothesis. Trends Neurosci. 33(1): 19.Google Scholar
Schiff, N. D. (2016). Central thalamic deep brain stimulation to support anterior forebrain mesocircuit function in the severely injured brain. J Neural Transmission (Vienna). 123(7): 797806.Google Scholar
Schiff, N. D., Plum, F. (2000). The role of arousal and “gating” systems in the neurology of impaired consciousness. J Clin Neurophysiol. 17(5): 438–52.Google Scholar
Schiff, N. D., et al. (2005). fMRI reveals large-scale network activation in minimally conscious patients. Neurology. 64: 514–23.Google Scholar
Schiff, N. D. (2007). Behavioural improvements with thalamic stimulation after severe traumatic brain injury. Nature. 448: 600–5.Google Scholar
Schiff, N. D., (2013). Gating of attentional effort through the central thalamus. J Neurophysiol. 109(4): 1152–63.Google Scholar
Schlosbeg, H. (1954). Three dimensions of emotion. Psychol Rev. 61: 81–8.Google Scholar
Schmahmann, J. D. (2003). Vascular syndromes of the thalamus. Stroke. 34(9): 2264–78.Google Scholar
Schmitt, I., et al. (2017). Thalamic amplification of cortical connectivity sustains attentional control. Nature. 545: 219–23.Google Scholar
Schober, J. M., Pfaff, D. (2007). The neurophysiology of sexual arousal. Best Pract Res Clin Endocrinol Metab. 21(3): 445–61.Google Scholar
Schober, J., Weil, Z., Pfaff, D. (2011). How generalized CNS arousal strengthens sexual arousal (and vice versa). Horm Behav. 59: 689–96.Google Scholar
Schölvinck, M. L., et al. (2010). Neural basis of global resting-state fMRI activity. Proc Natl Acad Sci U S A. 107(22): 10238–43.Google Scholar
Schultz, K. N., von Esenwein, S. A., Hu, M., et al. (2009) Article I. Viral vector-mediated overexpression of estrogen receptor-alpha in striatum enhances the estradiol-induced motor activity in female rats and estradiol-modulated GABA release. J Neurosci. 29(6): 1897–903. doi: 10.1523/JNEUROSCI.4647-08.2009.Google Scholar
Schultz, W. (2015). Neuronal reward and decision signals: from theories to data. Physiol Rev. 95(3): 853951.Google Scholar
Schultz, W. (2016a). Dopamine reward prediction error coding. Dialogues Clin Neurosci. 18(1): 2332.Google Scholar
Schultz, W. (2016b). Dopamine reward prediction-error signalling: a two-component response. Nat Rev Neurosci. 17(3): 183–95.Google Scholar
Schwabe, L., et al. (2013). Opposite effects of noradrenergic arousal on amygdala processing of fearful faces in men and women. Neuroimage. 73: 17.Google Scholar
Schwartz, J. C. (2011). The histamine H3 receptor: from discovery to clinical trials with pitolisant. Br J Pharmacol. 163(4): 713–21.Google Scholar
Scott, J. P. and Fuller, J. L. (1965). Genetics and the Social Behavior of the Dog. Chicago, IL: University of Chicago Press.Google Scholar
Segarra, A. C., et al. (2010). Estradiol: a key biological substrate mediating the response to cocaine in female rats. Horm Behav. 58(1): 3343.Google Scholar
Sehgal, A., Mignot, E. (2011). Genetics of sleep and sleep disorders. Cell. 146: 194207.Google Scholar
Senger, D. R., Claffey, K. P., Benes, J. E., et al. (1997). Angiogenesis promoted by vascular endothelial growth factor: regulation through alpha1beta1 and alpha2beta1 integrins. Proc Natl Acad Sci U S A. 94: 13612–17.Google Scholar
Senger, D. R., Perruzzi, C. A., Streit, M., et al. (2002). The alpha(1)beta(1) and alpha(2)beta(1) integrins provide critical support for vascular endothelial growth factor signaling, endothelial cell migration, and tumor angiogenesis. Am J Pathol. 160: 195204.Google Scholar
Serova, L., Rivkin, M., Nakashima, A., Sabban, E. L. (2002). Estradiol stimulates gene expression of norepinephrine biosynthetic enzymes in rat locus coeruleus. Neuroendocrinology. 75(3): 193200.Google Scholar
Serova, L. I., Maharjan, S., Sabban, E. L. (2005). Estrogen modifies stress response of catecholamine biosynthetic enzyme genes and cardiovascular system in ovariectomized female rats. Neuroscience. 132(2): 249–59.Google Scholar
Shang, Y., Griffith, L. C., Rosbash, M. (2008). Light-arousal and circadian photoreception circuits intersect at the large PDF cells of the Drosophila brain. Proc Natl Acad Sci U S A. 105(50): 19587–94.Google Scholar
Shannon, C. (1948). A mathematical theory of communication. Bell System Technical Journal. 27: 379397.Google Scholar
Sherin, J. E., Elmquist, J. K., Torrealba, F., Saper, C. B. (1998). Innervation of histaminergic tuberomammillary neurons by GABAergic and galaninergic neurons in the ventrolateral preoptic nucleus of the rat. J Neurosci. 18: 4705–21.Google Scholar
Sherin, J. E., Shiromani, P. J., McCarley, R. W., Saper, C. B. (1996). Activation of ventrolateral preoptic neurons during sleep. Science. 271(5246): 216–9.Google Scholar
Shigemori, M., et al. (1992). Coexisting diffuse axonal injury (DAI) and outcome of severe head injury. Acta Neurochir Suppl (Wien). 55: 37–9.Google Scholar
Shirvalkar, P., et al. (2006). Cognitive enhancement with central thalamic electrical stimulation. PNAS. 103: 17007–12.Google Scholar
Shouse, M. N., Siegel, J. M. (1992). Pontine regulation of REM sleep components in cats: integrity of the pedunculopontine tegmentum (PPT) is important for phasic events but unnecessary for atonia during REM sleep. Brain Res. 571(1): 5063.Google Scholar
Shulman, G. L. et al. (1997). Common blood flow changes across visual tasks: decreases in cerebral cortex. J Cogn Neurosci. 95: 648–63.Google Scholar
Shulman, R. G., Hyder, F., Rothman, D. L. (2009). Baseline brain energy supports the state of consciousness. Proc Natl Acad Sci U S A. 106(27): 11096–101.Google Scholar
Sieck, G. C., Harper, R. M. (1980). Discharge of neurons in the parabrachial pons related to the cardiac cycle: changes during different sleep-waking states. Brain Res. 199: 385–99.Google Scholar
Siegel, J. (2004). Brain mechanisms that control sleep and waking. Naturwissenschaften. 8: 355–65.Google Scholar
Silver, R., LeSauter, J., Tresco, P. A., Lehman, M. N. (1996) Article II. A diffusible coupling signal from the transplanted suprachiasmatic nucleus controlling circadian locomotor rhythms. Nature. 382(6594): 810–3.Google Scholar
Simon, H., Le Moal, M., Calas, A. (1979). Efferents and afferents of the ventral tegmental-Al0 region studied after local injection of [3H] leucine and horseradish peroxidase. Brain Res. 178: 1740.Google Scholar
Sinton, C. M., McCarley, R. W. (2004). Neurophysiological mechanisms of sleep and wakefulness: a question of balance. Semin Neurol. 24: 211–23.Google Scholar
Smith, H. R., Pang, K. C. (2005). Orexin-saporin lesions of the medial septum impair spatial memory. Neuroscience. 132(2): 261–71.Google Scholar
Smith, A. C., et al. (2009). A Bayesian statistical analysis of behavioral facilitation associated with deep brain stimulation. J Neurosci Meth. 183: 267–76.Google Scholar
Smith, R., Thayer, J. F., et al. (2017). The hierarchical basis of neurovisceral integration. Neurosci Biobehav Rev. 75: 274–96.Google Scholar
Smyser, C. D., et al. (2010). Longitudinal analysis of neural network development in preterm infants. Cereb Cortex. 20: 2852–62.Google Scholar
Snyder, A. Z. (2016). Intrinsic brain activity and resting state networks. In: Pfaff, D. W., Volkow, N. D.) Neuroscience in the 21st Century (2nd edition, volume 3). New York, NY: Springer Verlag, pp. 1626–83.Google Scholar
Solovey, G., et al. (2012). Self-regulated dynamical criticality in human ECoG. Front Integr Neurosci. 6: 44.Google Scholar
Solovey, G., (2015). Loss of consciousness is associated with stabilization of cortical activity. J Neurosci. 35(30): 10866–77.Google Scholar
Spiteri, T., et al. (2010). The role of the estrogen receptor alpha in the medial amygdala and ventromedial nucleus of the hypothalamus in social recognition, anxiety and aggression. Behav Brain Res. 210(2): 211–20.Google Scholar
Spiteri, T., Ogawa, S., Musatov, S., Pfaff, D. W., Agmo, A. (2012). The role of the estrogen receptor α in the medial preoptic area in sexual incentive motivation, proceptivity and receptivity, anxiety, and wheel running in female rats. Behav Brain Res. 230(1): 1120.Google Scholar
Sporns, O., Tononi, G., Edelman, G. M. (2000) Connectivity and complexity: the relationship between neuroanatomy and brain dynamics. Neural Netw. 13(8–9): 909–22.Google Scholar
Sporns, O. (2011). Networks of the Brain. Cambridge: MIT Press.Google Scholar
Squire, L. R., et al. (Eds.), Fundamental Neuroscience. San Diego, CA: Academic Press (Elsevier).Google Scholar
Stallings, M. C., et al. (1996). Genetic and environmental structure of the Tridimensional Personality Questionnaire: three or four temperament dimensions? J Pers Social Psychol. 70(1): 127–40.Google Scholar
Stam, C. J. (2014). Modern network science of neurological disorders. Nat Rev Neurosci. 15(10): 683–95.Google Scholar
Stanski, D. R., Shafer, S. (2005). Measuring depth of anesthesia. In: Miller, R. D. (Ed.), Miller's Anesthesia (6th edition). Philadelphia, PA: Churchill-Livingstone, pp. 1227–64.Google Scholar
Starzl, T. E., Taylor, C. W., Magoun, H. W. (1951). Ascending conduction in reticular activating system, with special reference to the diencephalon. J Neurophysiol. 14: 461–77.Google Scholar
Stauffer, W. R., Lak, A., Kobayashi, S., Schultz, W. (2016). Components and characteristics of the dopamine reward utility signal. J Comp Neurol. 524(8): 1699–711.Google Scholar
Steriade, M., (2003). The corticothalamic system in sleep. Front Biosci. 8: d878899.Google Scholar
Steriade, M., Paré, D., Bouhassira, D., Deschênes, M., Oakson, G. (1989). Phasic activation of lateral geniculate and perigeniculate thalamic neurons during sleep with ponto-geniculo-occipital waves. J Neurosci. 9(7): 2215–29.Google Scholar
Steriade, M., Datta, S., Pare, D., Oakson, G., Curro Dossi, R. C. (1990). Neuronal activation in brain-stem cholinergic nuclei related to tonic activation processes in thalamocortical systems. J Neurosci. 10: 2541–59.Google Scholar
Steriade, M., Dossi, R. C., Pare, D., Oakson, G. (1991). Fast oscillations (20–40 Hz) in thalamocortical systems and their potentiation by mesopontine cholinergic nuclei in the cat. Proc Natl Acad Sci. 88: 4396–400.Google Scholar
Steriade, M., McCormick, D. A., Sejnowski, T. J. (1993). Thalamocortical oscillations in the sleeping and aroused brain. Science. 262: 679–85.Google Scholar
Stevens, J. S. et al. (2017). Amygdala reactivity and anterior cingulated habituation predict posttraumatic stress disorder syndrome maintenance after acute civilian trauma. Biol Psychiatry. 81: 1023–9.Google Scholar
Studer, L. (2012). Derivation of dopaminergic neurons from pluripotent stem cells. Prog Brain Res. 200: 243–63.Google Scholar
Studer, L. (2017). Strategies for bringing stem cell-derived dopamine neurons to the clinic-The NYSTEM trial. Prog Brain Res. 230: 191212.Google Scholar
Sundermann, E. E., Maki, P. M., Bishop, J. R. (2010). A review of estrogen receptor alpha gene (ESR1) polymorphisms, mood, and cognition. Menopause. 17(4): 874–86.Google Scholar
Suzuki, D. T. (1960). Introduction to Zen Buddhism. London: Rider & Co.Google Scholar
Swanson, L. W. (1982). The projections of the ventral tegmental area and adjacent regions: a combined fluorescent retrograde tracer and immunofluorescence study in the rat. Brain Res Bull. 9: 321–53.Google Scholar
Swanson, L. W., Hartman, B. K. (1974). The central adrenergic system. J Comp Neurol. 163: 467506.Google Scholar
Tabansky, I., Quinkert, A. W., Rahman, N., et al. (2014). Temporally-patterned deep brain stimulation in a mouse model of multiple traumatic brain injury. Behav Brain Res. 273: 123–32.Google Scholar
Tabansky, I., Stern, J. N., Pfaff, D. W. (2015). Implications of epigenetic variability within a cell population for “Cell Type” classification. Front Behav Neurosci. 9: 342.Google Scholar
Tabansky, I., et al. (2018). Molecular profiling of reticular gigantocellularis neurons indicates that eNOS modulates environmentally dependent levels of arousal. Proc Natl Acad Sci. 115: 69006909. doi: 10.1073/pnas.1806123115.Google Scholar
Taheri, S., Zeitzer, J. M., Mignot, E. (2002). The role of hypocretins (orexins) in sleep regulation and narcolepsy. Ann Rev Neurosci. 25: 283313.Google Scholar
Takahashi, J. S. (2016). Molecular architecture of the circadian clock in mammals. In: Sassone-Corsi, P., Christen, Y., (Eds.), A Time for Metabolism and Hormones [Internet]. Cham (CH): Springer.Google Scholar
Takamori, S., Rhee, J. S., Rosenmund, C., Jahn, R. (2001). Identification of differentiation-associated brain-specific phosphate transporter as a second vesicular glutamate transporter (VGLUT2). J Neurosci. 21: RC182.Google Scholar
Takigawa, M., Mogenson, G. J. (1977). A study of inputs to antidromically identified neurons of the locus coeruleus. Brain Res. 135: 217–30.Google Scholar
Tan, K., Le Douarin, N. M. (1991). Development of the nuclei and cell migration in the medulla oblongata. Application of the quail-chick chimera system. Anat Embryol (Berl). 183(4): 321–43.Google Scholar
Tang, W., et al. (2017). Dynamic connectivity modulates local activity in the core regions of the default-mode network. PNAS. 114: 9713–18.Google Scholar
Taylor, P. N., Wang, Y., Kaiser, M. (2017). Within brain area tractography suggests local modularity using high resolution connectomics. Sci Rep. 7: 39859. doi: 10.1038/srep39859.Google Scholar
Thibaut, A., et al. (2017). Controlled clinical trial of repeated prefrontal tDCS in patients with chronic minimally conscious state. Brain Inj. 31(4): 466–74.Google Scholar
Thompson, R. H., Swanson, L. W. (2003). Structural characterization of a hypothalamic visceromotor pattern generator network. Brain Res Rev. 41: 153202.Google Scholar
Thor, D. H., Wainwright, K. L., Holloway, W. R. (1982). Persistence of attention to a novel conspecific: some developmental variables in laboratory rats. Dev Psychobiol. 15(1): 18.Google Scholar
Tong, L. (2005). Acetyl-coenzyme A carboxylase: crucial metabolic enzyme and attractive target for drug discovery. Cell Mol Life Sci. 62: 1784–803.Google Scholar
Tonge, D. A., de Burgh, H. T., Docherty R., et al. (2012). Fibronectin supports neurite outgrowth and axonal regeneration of adult brain neurons in vitro. Brain Res. 1453: 816.Google Scholar
Tononi, G., Edelman, G. M. (1998) Consciousness and complexity. Science. 282(5395): 1846–51.Google Scholar
Top, D., Young, M. W. (2017). Coordination between differentially regulated circadian clocks generates rhythmic behavior. Cold Spring Harb Perspect Biol. pii:a033589. doi: 10.1101/cshperspect.a033589.Google Scholar
Trofimova, I., Robbins, T. W. (2016). Temperament and arousal systems: a new synthesis of differential psychology and functional neurochemistry. Neurosci Biobehav Rev. 64: 382402.Google Scholar
Tsunematsu, T., Kilduff, T. S., Boyden, E. S. et al. (2011) Article II. Acute optogenetic silencing of orexin/hypocretin neurons induces slow-wave sleep in mice. J Neurosci. 31(29): 10529–39. doi: 10.1523/JNEUROSCI.0784-11.2011.Google Scholar
Turney, S. G., Bridgman, P. C. (2005). Laminin stimulates and guides axonal outgrowth via growth cone myosin II activity. Nat Neurosci. 8: 717–19.Google Scholar
Valentino, R. J., Foote, S. L. (1988). Corticotropin-releasing hormone increases tonic but not sensory-evoked activity of noradrenergic locus coeruleus neurons in unanesthetized rats. J Neurosci. 8(3): 1016–25.Google Scholar
Valentino, R. J., Foote, S. L., Aston-Jones, G. (1983). Corticotropin-releasing factor activates noradrenergic neurons of the locus coeruleus. Brain Res. 270(2): 363–7.Google Scholar
Valentino, R. J., Page, M., Van Bockstaele, E., Aston-Jones, G. (1992). Corticotropin-releasing factor innervation of the locus coeruleus region: distribution of fibers and sources of input. Neuroscience. 48: 689705.Google Scholar
Valentino, R. J., Rudoy, C., Saunders, A., Liu, X. B., Van Bockstaele, E. J. (2001). Corticotropin-releasing factor is preferentially colocalized with excitatory rather than inhibitory amino acids in axon terminals in the peri-locus coeruleus region. Neuroscience. 106(2): 375–84.Google Scholar
Valverde, F. (1961). Reticular formation of the pons and medulla oblongata; a Golgi study. J Comp Neurol. 116: 7199.Google Scholar
Valverde, F. (1962). Reticular formation of the albino rat's brain stem. J Comp Neurol. 119: 2549.Google Scholar
van Someren, E., Cluydts R. (2017). Sleep regulation and insomnia. In Pfaff, D. W., Volkow, N. D., Neuroscience in the 21st Century (2nd edition, volume 3). Heidelberg, New York: Springer, pp. 2289–316.Google Scholar
van Swinderen, B., Greenspan, R. J. (2003). Salience modulates 20–30 Hz brain activity in Drosophila. Nat Neurosci. 6(6): 579–86.Google Scholar
Vanderwolf, C. H., Stewart, D. J. (1998). Thalamic control of neocortical activation: a critical re-evaluation. Brain Res Bull. 20: 529–38.Google Scholar
Vasquez-Vivar, J., Kalyanaraman, B., Martasek, P., et al. (1998). Superoxide generation by endothelial nitric oxide synthase: the influence of cofactors. Proc Natl Acad Sci U S A. 95: 9220–5.Google Scholar
Vazey, E. M., Aston-Jones, G. (2014). Designer receptor manipulations reveal a role of the locus coeruleus noradrenergic system in isoflurane general anesthesia. Proc Natl Acad Sci U S A. 111(10): 3859–64.Google Scholar
Vertes, R. P., Martin, G. F., Waltzer, R. (1986). An autoradiographic analysis of ascending projections from the medullary reticular formation in the rat. Neuroscience. 19: 873–98.Google Scholar
Villablanca, J., Salinas-Zeballos, M. E. (1972). Sleep-wakefulness, EEG and behavioral studies of chronic cats without the thalamus: the ‘athalamic’ cat. Arch Ital Biol. 110: 383411.Google Scholar
Villano, I., Messina, A., Valenzano, A. et al. (2017). Basal forebrain cholinergic system and orexin neurons: effects on attention. Front Behav Neurosci. 11: 10. doi: 10.3389/fnbeh.2017.00010.Google Scholar
Vincent, J. L., et al. (2007). Intrinsic functional architecture in the anesthetized monkey brain. Nature. 447: 46–7.Google Scholar
Von Economo, J. (1926). Handbuch der Normalen und Pathologischen Physiologie. Berlin: Springer, pp. 591610.Google Scholar
Vosshall, L. B, et al. (1994). Block in nuclear localization of period protein by a second clock mutation, timeless. Science. 263(5153): 1606–9.Google Scholar
Vosshall, L. B., Young, M. W. (1995). Circadian rhythms in Drosophila can be driven by period expression in a restricted group of central brain cells. Neuron. 15(2): 345–60.Google Scholar
Vujovic, N., et al. (2015). Projections from the subparaventricular zone define four channels of output from the circadian timing system. J Comp Neurol. 523(18): 2714–37.Google Scholar
Waid, D. K., Chell, M., El-Fakahany, E. E. (2000). M(2) and M(4) muscarinic receptor subtypes couple to activation of endothelial nitric oxide synthase. Pharmacology. 61: 3742.Google Scholar
Wang, C., et al. (1996). Testosterone replacement therapy improves mood in hypogonadal men – a clinical research center study. J Clin Endocrinol Metab. 81(10): 3578–83.Google Scholar
Wang, S., et al. (2017). The human amygdala parametrically encodes the intensity of specific facial emotions and their categorical ambiguity. Nat Commun. 21(8): 14821. doi: 10.1038/ncomms14821.Google Scholar
Wannez, S., et al. (2017). Prevalence of coma-recovery scale-revised signs of consciousness in patients in minimally conscious state. Neuropsychol Rehabil. 11: 110.Google Scholar
Watson, C. J., Lydic, R., Baghdoyan, H. A. (2011). Sleep duration varies as a function of glutamate and GABA in rat pontine reticular formation. J Neurochem. 118(4): 571–80.Google Scholar
Weber, F., Yan, Y. (2016). Circuit-based interrogation of sleep control. Nature. 538: 5161.Google Scholar
Webster, H. H., Jones, B. E. (1988). Neurotoxic lesions of the dorsolateral pontomesencephalic tegmentum-cholinergic cell area in the cat. II. Effects upon sleep-waking states. Brain Res. 458(2): 285302.Google Scholar
Weil, Z. M., et al. (2010). Impact of generalized brain arousal on sexual behavior. Proc Natl Acad Sci. 107(5): 2265–70.Google Scholar
Weinberger, J. (2018). A Brief History of Unconscious Processes. New Haven, CT: Guilford PressGoogle Scholar
Weitzman, E. D. (1981). Sleep and its disorders. Annu Rev Neurosci. 4: 381417.Google Scholar
Wenk, G. L., Stoehr, J. D., Quintana, G., Mobley, S., Wiley, R. G. (1994). Behavioral, biochemical, histological, and electrophysiological effects of 192 IgG-saporin injections into the basal forebrain of rats. J Neurosci. 14: 5986–95.Google Scholar
Williams, K. M., Mong, J. A. (2017). Methamphetamine and ovarian steroid responsive cells in the posteriodorsal medial amygdala are required for methamphetamine-enhanced proceptive behaviors. Science Rep. 2017 7: 39817.Google Scholar
Winsky-Sommerer, R., Yamanaka, A., Diano, S., et al., (2004). Interaction between the corticotrophin-releasing factor system and hypocretins (orexins): a novel circuit mediating stress response. J Neurosci. 24: 11439–48.Google Scholar
Wood, S. K., Valentino, R. J. (2017). The brain norepinephrine system, stress and cardiovascular vulnerability. Neurosci Biobehav Rev. 74(Pt B): 393400.Google Scholar
Wu, H. B., Stavarache, M., Pfaff, D. W., Kow, L. M. (2007). Arousal of cerebral cortex electroencephalogram consequent to high-frequency stimulation of ventral medullary reticular formation. Proc Natl Acad Sci U S A. 104(46): 18292–6.Google Scholar
Wu, H., Stavarache, M., Pfaff, D. W., Kow, L. (2007). Arousal of cerebral cortex electroencephalogram consequent to high-frequency stimulation of ventral medullary reticular formation. Proc Natl Acad Sci. 104(46): 18292–6.Google Scholar
Xia, Y., Tsai, A. L., Berka, V., Zweier, J. L. (1998). Superoxide generation from endothelial nitric-oxide synthase. A Ca2+/calmodulin-dependent and tetrahydrobiopterin regulatory process. J Biol Chem. 273: 25804–8.Google Scholar
Xu, C., Datta, S., Wu, M., Alreja, M. (2004). Hippocampal theta rhythm is reduced by suppression of the H-current in septohippocampal GABAergic neurons. Eur J Neurosci. 19: 2299–309.Google Scholar
Yackle, K., et al. (2017). Breathing control center neurons that promote arousal in mice. Science. 355: 1411–15.Google Scholar
Yamanaka, A. et al. (2003). Hypothalamic orexin neurons regulate arousal according to energy balance in mice. Neuron. 38(5): 701–13.Google Scholar
Yanofsky, N. (2016). Paradoxes, contradictions and the limits of science. Sci Am. 104: 166–78.Google Scholar
Yokota, S., Oka, T., Tsumori, T., Nakamura, S., Yasui, Y. (2007). Glutamatergic neurons in the Kolliker-Fuse nucleus project to the rostral ventral respiratory group and phrenic nucleus. A combined retrograde tracing and in situ hybridization study in the rat. Neurosci Res. 59: 342–6.Google Scholar
Yokota, S., et al. (2015). Respiratory-related outputs of glutamatergic, hypercapnia-responsive parabrachial neurons in mice. J Comp Neurol. 523(6): 907–20.Google Scholar
Yoshida, Y., et al. (2001). Fluctuation of extracellular hypocretin-1 (orexin A) levels in the rat in relation to the light-dark cycle and sleep-wake activities. Eur J Neurosci. 14(7): 1075–81.Google Scholar
Young, J. W. (2009). Dopamine D1 and D2 receptor family contributions to modafinil-induced wakefulness. J Neurosci. 29: 2663–5.Google Scholar
Young, M. W. (2002). Big ben rings in a lesson on biological clocks. Neuron. 36: 1001–5.Google Scholar
Young, M. W., Kay, S. A. (2001). Time zones: a comparative genetics of circadian clocks. Nat Rev Genet. 2: 702–15.Google Scholar
Zaborszky, L., Cullinan, W. E. (1989). Hypothalamic axons terminate on forebrain cholinergic neurons: an ultrastructural double-labeling study using PHA-L tracing and ChAT immunocytochemistry. Brain Res. 479: 177–84.Google Scholar
Zaborszky, L., et al. (2005). Three-dimensional chemoarchitecture of the basal forebrain: Spatially specific association of cholinergic and calcium binding protein-containing neurons. Neuroscience. 136: 697713.Google Scholar
Zaborszky, L., Pol, A. V. D., Gyengesi, E., (2012). The basal forebrain cholinergic projection system in mice. In: Watson, C., Paxinos, G., Puelles, L. (Eds.), The Mouse Nervous System (1st edition). Amsterdam: Elsevier, pp. 684718.Google Scholar
Zaborszky, L., et al. (2015). Neurons in the basal forebrain project to the cortex in a complex topographic organization that reflects corticocortical connectivity patterns: an experimental study based on retrograde tracing and 3D reconstruction. Cerebr Cortex. 25: 118–37.Google Scholar
Zagha, E., McCormick, D. A. (2014). Neural control of brain state. Curr Opin Neurobiol. 29: 178–86.Google Scholar
Zavalko, I. et al. (2012). Hypersomnia due to bilateral thalamic lesions: unexpected response to Modafinil. Eur J Neurol. 19: 125–37.Google Scholar
Zeitzer, J. M., Nishino, S., Mignot, E. (2006). The neurobiology of hypocretins (orexins), narcolepsy and related therapeutic interventions. Trends Pharmacol Sci. 27(7): 368–74.Google Scholar
Zeman, A. (2001). Consciousness Brain. 124: 1263–89.Google Scholar
Zeman, A. (2002). Consciousness, a User's Guide. New Haven, CT: Yale University Press.Google Scholar
Zemlan, F. P., Kow, L. M., Pfaff, D. W. (1983). Effect of interruption of bulbospinal pathways on lordosis, posture, and locomotion. Exp Neurol. 81(1): 177–94.Google Scholar
Zemlan, F. P., Behbehani, M. M., Beckstead, R. M. (1984). Ascending and descending projections from nucleus reticularis magnocellularis and nucleus reticularis gigantocellularis: an autoradiographic and horseradish peroxidase study in the rat. Brain Res. 292: 207–20.Google Scholar
Zhao, W., Becker, J. B. (2010). Sensitization enhances acquisition of cocaine self-administration in female rats: estradiol further enhances cocaine intake after acquisition. Horm Behav. 58(1): 812.Google Scholar
Zhao, X., et al. (2014). Nuclear receptors rock around the clock. EMBO Rep. 15(5): 518–28.Google Scholar
Zhou, J., et al. (2009). Arousal-related reticular neurons during reduced oxygen tension: resilience and recovery of electrical activity. Dev Neurosci. 31(4): 255–8.Google Scholar
Zoubina, E. V., Smith, P. G. (2003). Expression of estrogen receptors alpha and beta by sympathetic ganglion neurons projecting to the proximal urethra of female rats. J Urol. 169: 382–5.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • Bibliography
  • Donald Pfaff
  • Book: How Brain Arousal Mechanisms Work
  • Online publication: 29 November 2018
  • Chapter DOI: https://doi.org/10.1017/9781108377485.012
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.

  • Bibliography
  • Donald Pfaff
  • Book: How Brain Arousal Mechanisms Work
  • Online publication: 29 November 2018
  • Chapter DOI: https://doi.org/10.1017/9781108377485.012
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.

  • Bibliography
  • Donald Pfaff
  • Book: How Brain Arousal Mechanisms Work
  • Online publication: 29 November 2018
  • Chapter DOI: https://doi.org/10.1017/9781108377485.012
Available formats
×