Skip to main content Accessibility help
×
Hostname: page-component-7bb8b95d7b-nptnm Total loading time: 0 Render date: 2024-09-23T11:39:36.299Z Has data issue: false hasContentIssue false

Chapter 30 - Sudden Infant Death Syndrome from the Brainstem Perspective

from Section 7 - Pathophysiology

Published online by Cambridge University Press:  04 June 2019

Marta C. Cohen
Affiliation:
Sheffield Children’s Hospital
Irene B. Scheimberg
Affiliation:
Royal London Hospital
J. Bruce Beckwith
Affiliation:
Loma Linda University School of Medicine
Fern R. Hauck
Affiliation:
University of Virginia
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
Publisher: Cambridge University Press
Print publication year: 2019

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

Chiodini, BA, Thach, BT. Impaired ventilation in infants sleeping facedown: potential significance for Sudden Infant Death Syndrome. J Pediatr, 1993; 123:686–92.CrossRefGoogle ScholarPubMed
de Jonge, GA, Engelberts, AC, Koomen-Liefting, AJ, Kostense, PJ. Cot death and prone sleeping position in the Netherlands. BMJ, 1989; 298:722.CrossRefGoogle ScholarPubMed
Kemp, JS, Kowalski, RM, Burch, PM, Graham, MA, Thach, BT. Unintentional suffocation by rebreathing: a death scene and physiologic investigation of a possible cause of sudden infant death. J Pediatr, 1993; 122:874–80.CrossRefGoogle ScholarPubMed
McGlashan, ND. Sudden infant deaths in Tasmania, 1980–1986: a seven-year prospective study. Soc Sci Med, 1989; 29:1015–26.Google Scholar
Trachtenberg, FL, Haas, EA, Kinney, HC, Stanley, C, Krous, HF. Risk factor changes for Sudden Infant Death Syndrome after initiation of Back-to-Sleep campaign. Pediatrics, 2012; 129:630–8.Google Scholar
Cerpa, VJ, Aylwin, ML Beltrán-Castillo, S, Bravo, EU, Llona, IS et al. The alteration of neonatal raphe neurons by prenatal–perinatal nicotine meaning for Sudden Infant Death Syndrome. Am J Respir Cell Mol Biol, 2015; 53:489–99.Google Scholar
Leung, CG, Mason, P. Physiological properties of raphe magnus neurons during sleep and waking. J Neurophysiol, 1999; 81:584–95.CrossRefGoogle ScholarPubMed
Bérard, A, Zhao, JP, Sheehy, O. Success of smoking cessation interventions during pregnancy. Am J Obstet Gynecol, 2016; 215:611.CrossRefGoogle ScholarPubMed
Mage, DT, Donner, M. A unifying theory for SIDS. Int J Pediatr, 2009: 9:368270.Google Scholar
Paterson, DS, Thompson, EG, Kinney, HC. Serotonergic and glutamatergic neurons at the ventral medullary surface of the human infant: Observations relevant to central chemosensitivity in early human life. Auton Neurosci, 2006a; 124:112–24.CrossRefGoogle ScholarPubMed
Paterson, DS, Trachtenberg, FL, Thompson, EG, Belliveau, RA, Beggs, AH, Darnall, R, et al. Multiple serotonergic brainstem abnormalities in Sudden Infant Death Syndrome. JAMA, 2006b; 296:2124–32.Google Scholar
Carlin, RF, Moon, RY. Risk factors, protective factors, and current recommendations to reduce Sudden Infant Death Syndrome: a review. JAMA Pediatr, 2017; 171, 175–80.Google Scholar
Van Norstrand, DW, Ackerman, MJ. Genomic risk factors in Sudden Infant Death Syndrome. Genome Med, 2010; 2:86.CrossRefGoogle ScholarPubMed
Filiano, JJ, Kinney, HC. A perspective on neuropathologic findings in victims of the Sudden Infant Death Syndrome: the triple-risk model. Biol Neonate, 1994; 65:194–7.CrossRefGoogle ScholarPubMed
Kinney, HC, Thach, BT. The Sudden Infant Death Syndrome. N Engl J Med, 2009; 361:795805.Google Scholar
Glogowska, M, Richardson, PS, Widdicombe, JG, Winning, AJ. The role of the vagus nerves, peripheral chemoreceptors, and other afferent pathways in the genesis of augmented breaths in cats and rabbits. Respir Physiol, 1972; 16:179–96.CrossRefGoogle ScholarPubMed
McGinty, DJ, London, MS, Baker, TL, Stevenson, M, Hoppenbrouwers, T, Harper, RM, et al. Sleep apnea in normal kittens. Sleep, 1979; 1:393412.Google ScholarPubMed
Orem, J, Trotter, RH. Medullary respiratory neuronal activity during augmented breaths in intact unanesthetized cats. J Appl Physiol (1985), 1993; 74:761–9.Google Scholar
Lijowska, AS, Reed, NW, Chiodini, BA, Thach, BT. Sequential arousal and airway-defensive behavior of infants in asphyxial sleep environments. J Appl Physiol (1985), 1997; 83:219–28.CrossRefGoogle ScholarPubMed
McNamara, F, Wulbrand, H, Thach, BT. Characteristics of the infant arousal response. J Appl Physiol (1985), 1998; 85:2314–21.CrossRefGoogle ScholarPubMed
Thach, BT, Lijowska, A. Arousals in infants. Sleep, 1996; 19:S271–3.Google Scholar
Anderson, CA, Dick, TE, Orem, J. Respiratory responses to tracheobronchial stimulation during sleep and wakefulness in the adult cat. Sleep, 1996; 19:472–8.CrossRefGoogle ScholarPubMed
Kahn, A, Blum, D, Rebuffat, E, Sottiaux, M, Levitt, J, Bochner, A, et al. Polysomnographic studies of infants who subsequently died of Sudden Infant Death Syndrome. Pediatrics, 1988; 82:721–7.CrossRefGoogle ScholarPubMed
Bartlett, D. Origin and regulation of spontaneous deep breaths. Respir Physiol, 1971; 12:230–8.Google Scholar
Bell, HJ, Haouzi, P. Acetazolamide suppresses the prevalence of augmented breaths during exposure to hypoxia. Am J Physiol Regul Integr Comp Physiol, 2009; 297:R370–81.Google Scholar
Cherniack, NS, von Euler, C, Glogowska, M, Homma, I. Characteristics and rate of occurrence of spontaneous and provoked augmented breaths. Acta Physiol Scand, 1981; 111:349–60.Google Scholar
Hill, AA, Garcia, AJ, Zanella, S, Upadhyaya, R, Ramirez, JM. Graded reductions in oxygenation evoke graded reconfiguration of the isolated respiratory network. J Neurophysiol, 2011: 105:625–39.Google Scholar
Lieske, SP, Thoby-Brisson, M, Telgkamp, P, Ramirez, JM. Reconfiguration of the neural network controlling multiple breathing patterns: eupnea, sighs and gasps. Nat Neurosci, 2000; 3:600–7.Google Scholar
Smith, JC, Ellenberger, HH, Ballanyi, K, Richter, DW, Feldman, JL. Pre-Bötzinger complex: a brainstem region that may generate respiratory rhythm in mammals. Science, 1991; 254:726–9.Google Scholar
Schwarzacher, SW, Rub, U, Deller, T. Neuroanatomical characteristics of the human pre-Bötzinger complex and its involvement in neurodegenerative brainstem diseases. Brain, 2011; 134:2435.CrossRefGoogle ScholarPubMed
Tan, W, Janczewski, WA, Yang, P, Shao, XM, Callaway, EM, Feldman, JL. Silencing preBötzinger complex somatostatin-expressing neurons induces persistent apnea in awake rat. Nat Neurosci, 2008; 11:538–40.Google Scholar
Burke, PG, Abbott, SB, Coates, MB, Viar, KE, Stornetta, RL, Guyenet, PG. Optogenetic stimulation of adrenergic C1 neurons causes sleep state-dependent cardiorespiratory stimulation and arousal with sighs in rats. Am J Respir Crit Care Med, 2014; 190:1301–10.CrossRefGoogle ScholarPubMed
Dunne, KP, Fox, GP, O’Regan, M, Matthews, TG. Arousal responses in babies at risk of Sudden Infant Death Syndrome at different postnatal ages. Ir Med J, 1992; 85:1922.Google Scholar
Kahn, A, Groswasser, J, Rebuffat, E, Sottiaux, M, Blum, D, Foerster, M, et al. Sleep and cardiorespiratory characteristics of infant victims of sudden death: a prospective case-control study. Sleep, 1992; 15:287–92.Google Scholar
Kato, I, Scaillet, S, Groswasser, J, Montemitro, E, Togari, H, Lin, JS, et al. Spontaneous arousability in prone and supine position in healthy infants. Sleep, 2006; 29:785–90.CrossRefGoogle ScholarPubMed
McCulloch, K, Brouillette, RT, Guzzetta, AJ, Hunt, CE. Arousal responses in near-miss Sudden Infant Death Syndrome and in normal infants. J Pediatr, 1982; 101:911–17.Google Scholar
Sawaguchi, T, Kato, I, Franco, P, Sottiaux, M, Kadhim, H, Shimizu, S, et al. Apnea, glial apoptosis, and neuronal plasticity in the arousal pathway of victims of SIDS. Forensic Sci Int, 2005; 149:205–17.Google Scholar
Schechtman, VL, Harper, RM, Wilson, AJ, Southall, DP. Sleep state organization in normal infants and victims of the Sudden Infant Death Syndrome. Pediatrics, 1992; 89:865–70.Google Scholar
Bamford, OS, Schuen, JN, Carroll, JL. Effect of nicotine exposure on postnatal ventilatory responses to hypoxia and hypercapnia. Respir Physiol, 1996; 106:111.Google Scholar
Horne, RS, Sly, DJ, Cranage, SM, Chau, B, Adamson, TM. Effects of prematurity on arousal from sleep in the newborn infant. Pediatr Res, 2000; 47:468–74.Google Scholar
Nock, ML, Difiore, JM, Arko, MK, Martin, RJ. Relationship of the ventilatory response to hypoxia with neonatal apnea in preterm infants. J Pediatr, 2004; 144:291–5.Google Scholar
Hehre, DA, Devia, CJ, Bancalari, E, Suguihara, C. Brainstem amino acid neurotransmitters and ventilatory response to hypoxia in piglets. Pediatr Res, 2008; 63:4650.Google Scholar
Horne, RS, Parslow, PM, Harding, R. Postnatal development of ventilatory and arousal responses to hypoxia in human infants. Respir Physiol Neurobiol, 2005; 149:257–71.Google Scholar
Haupt, ME, Goodman, DM, Sheldon, SH. Sleep-related expiratory obstructive apnea in children. J Clin Sleep Med, 2012; 8:673–9.CrossRefGoogle ScholarPubMed
Porges, WL, Hennessy, EJ, Quail, AW, Cottee, DB, Moore, PG, McIlveen, SA, et al. Heart-lung interactions: the sigh and autonomic control in the bronchial and coronary circulations. Clin Exp Pharmacol Physiol, 2000; 27:1022–7.CrossRefGoogle ScholarPubMed
Weese-Mayer, DE, Kenny, AS, Bennett, HL, Ramirez, JM, Leurgans, SE. Familial dysautonomia: frequent, prolonged, and severe hypoxemia during wakefulness and sleep. Pediatr Pulmonol, 2008; 43:251–60.Google Scholar
Wulbrand, H, McNamara, F, Thach, BT. The role of arousal related brainstem reflexes in causing recovery from upper airway occlusion in infants. Sleep, 2008; 31:833–40.Google Scholar
Thach, BT. Graded arousal responses in infants: advantages and disadvantages of a low threshold for arousal. Sleep Med, 2002; 3(Suppl 2):S37-40.Google Scholar
Franco, P, Szliwowski, H, Dramaix, M, Kahn, A. Polysomnographic study of the autonomic nervous system in potential victims of Sudden Infant Death Syndrome. Clin Auton Res, 1998; 8:243–9.Google Scholar
Franco, P, Verheulpen, D, Valente, F, Kelmanson, I, de Broca, A, Scaillet, S, et al. Autonomic responses to sighs in healthy infants and in victims of sudden infant death. Sleep Med, 2003; 4:569–77.CrossRefGoogle ScholarPubMed
Lucchini, M, Signorini, MG, Fifer, WP, Sahni, R Multi-parametric heart rate analysis in premature babies exposed to Sudden Infant Death Syndrome. Conf Proc IEEE Eng Med Biol Soc, 2014; 2014:6389–92; https://www.ncbi.nlm.nih.gov/pubmed/25571458 (accessed 31 October 2018).Google Scholar
Pena, F, Parkis, MA, Tryba, AK, Ramirez, JM. Differential contribution of pacemaker properties to the generation of respiratory rhythms during normoxia and hypoxia. Neuron, 2004; 43:105–17.CrossRefGoogle Scholar
Harper, RM, Kinney, HC, Fleming, PJ, Thach, BT. Sleep influences on homeostatic functions: implications for Sudden Infant Death Syndrome. Respir Physiol, 2000; 119:123–32.Google Scholar
Hunt, CE. The cardiorespiratory control hypothesis for Sudden Infant Death Syndrome. Clin Perinatol, 1992; 19:757–71.Google Scholar
Poets, CF, Meny, RG, Chobanian, MR, Bonofiglo, RE. Gasping and other cardiorespiratory patterns during sudden infant deaths. Pediatr Res, 1999; 45:350–4.CrossRefGoogle ScholarPubMed
Cherniack, NS, Edelman, NH, Lahiri, S. The effect of hypoxia and hypercapnia on respiratory neuron activity and cerebral aerobic metabolism. Chest, 1971; 59(Suppl):29S.Google Scholar
Pena, F, Aguileta, MA. Effects of riluzole and flufenamic acid on eupnea and gasping of neonatal mice in vivo. Neurosci Lett, 2007; 415:288–93.Google Scholar
Erickson, JT, Sposato, BC Autoresuscitation responses to hypoxia-induced apnea are delayed in newborn 5-HT-deficient Pet-1 homozygous mice. J Appl Physiol (1985). 2009; 106(6):1785–92.Google Scholar
Serdarevich, C, Fewell, JE. Influence of core temperature on autoresuscitation during repeated exposure to hypoxia in normal rat pups. J Appl Physiol (1985), 1999; 87:1346–53.Google Scholar
Ramirez, JM, Dashevskiy, T, Marlin, IA, Baertsch, N. Microcircuits in respiratory rhythm generation: commonalities with other rhythm generating networks and evolutionary perspectives. Curr Opin Neurobiol, 2016; 41:5361.Google Scholar
Anderson, TM, Garcia, AJ, Baertsch, NA, Pollak, J, Bloom, JC, Wei, AD, et al. A novel excitatory network for the control of breathing. Nature, 2016; 536:7680.Google Scholar
Janczewski, WA, Feldman, JL. Distinct rhythm generators for inspiration and expiration in the juvenile rat. J Physiol, 2006; 570:407–20.Google Scholar
Pagliardini, S, Janczewski, WA, Tan, W, Dickson, CT, Deisseroth, K, Feldman, JL. Active expiration induced by excitation of ventral medulla in adult anesthetized rats. J Neurosci, 2011; 31:2895–905.Google Scholar
Guyenet, PG. Regulation of breathing and autonomic outflows by chemoreceptors. Compr Physiol, 2014; 4:1511–62.Google Scholar
Guyenet, PG, Bayliss, DA. Neural control of breathing and CO2 homeostasis. Neuron, 2015; 87:946–61.CrossRefGoogle ScholarPubMed
Kumar, NN, Velic, A, Soliz, J, Shi, Y, Li, K, Wang, S, et al. PHYSIOLOGY. Regulation of breathing by CO(2) requires the proton-activated receptor GPR4 in retrotrapezoid nucleus neurons. Science, 2015; 348:1255–60.Google Scholar
Ramanantsoa, N, Hirsch, MR, Thoby-Brisson, M, Dubreuil, V, Bouvier, J, Ruffault, PL et al. Breathing without CO(2) chemosensitivity in conditional Phox2b mutants. J Neurosci, 2011; 31:12880–8.Google Scholar
Ruffault, PL, D’Autreaux, F, Hayes, JA, Nomaksteinsky, M, Autran, S, Fujiyama, T et al. The retrotrapezoid nucleus neurons expressing Atoh1 and Phox2b are essential for the respiratory response to CO(2). eLife, 2015; 4:e07051; https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4429526/ (accessed 31 October 2018).Google Scholar
Mendelowitz, D. Advances in parasympathetic control of heart rate and cardiac function. News Physiol Sci, 1999; 14:155–61.Google ScholarPubMed
Neff, RA, Simmens, SJ, Evans, C, Mendelowitz, D. Prenatal nicotine exposure alters central cardiorespiratory responses to hypoxia in rats: implications for Sudden Infant Death Syndrome. J Neurosci, 2004; 24:9261–8.Google Scholar
Carroll, MS, Kenny, AS, Patwari, PP, Ramirez, JM, Weese-Mayer, DE. Respiratory and cardiovascular indicators of autonomic nervous system dysregulation in familial dysautonomia. Pediatr Pulmonol, 2012; 47:682–91.Google Scholar
Garcia, AJ, Koschnitzky, JE, Dashevskiy, T, Ramirez, JM. Cardiorespiratory coupling in health and disease. Auton Neurosci, 2013; 175:2637.Google Scholar
Fu, W, Le Maitre, E, Fabre, V, Bernard, JF,David Xu, ZQ, Hokfelt, T. Chemical neuroanatomy of the dorsal raphe nucleus and adjacent structures of the mouse brain. J Comp Neurol, 2010 518:3464–94.Google Scholar
Stornetta, RL, Rosin, DL, Simmons, JR, McQuiston, TJ, Vujovic, N, Weston, MC, Guyenet, PG. Coexpression of vesicular glutamate transporter-3 and gamma-aminobutyric acidergic markers in rat rostral medullary raphe and intermediolateral cell column. J Comp Neurol, 2005; 492:477–94.Google Scholar
Accorsi-Mendonca, D, Castania, JA, Bonagamba, LG, Machado, BH, Leao, RM. Synaptic profile of nucleus tractus solitarius neurons involved with the peripheral chemoreflex pathways. Neuroscience, 2011; 197:107–20.Google Scholar
Chitravanshi, VC, Sapru, HN. Chemoreceptor-sensitive neurons in commissural subnucleus of nucleus tractus solitarius of the rat. Am J Physiol, 1995; 268:R851–8.Google Scholar
Machado, BH. Neurotransmission of the cardiovascular reflexes in the nucleus tractus solitarii of awake rats. Ann N Y Acad Sci, 2001; 940:179–96.Google Scholar
Mifflin, SW. Arterial chemoreceptor input to nucleus tractus solitarius. Am J Physiol, 1992; 263:R368–75.Google Scholar
Brannan, S, Liotti, M, Egan, G, Shade, R, Madden, L, Robillard, R, et al. Neuroimaging of cerebral activations and deactivations associated with hypercapnia and hunger for air. Proc Natl Acad Sci USA, 2001; 98:2029–34.Google Scholar
Burdakov, D, Karnani, MM, Gonzalez, A. Lateral hypothalamus as a sensor-regulator in respiratory and metabolic control. Physiol Behav, 2013; 121:117–24.CrossRefGoogle ScholarPubMed
Chamberlin, NL, Saper, CB. Topographic organization of respiratory responses to glutamate microstimulation of the parabrachial nucleus in the rat. J Neurosci, 1994; 14:6500–10.Google Scholar
Masaoka, Y, Sugiyama, H, Katayama, A, Kashiwagi, M, Homma, I. Slow breathing and emotions associated with odor-induced autobiographical memories. Chem Senses, 2012; 37:379–88.Google Scholar
Nattie, E, Li, A. Respiration and autonomic regulation and orexin. Prog Brain Res, 2012; 198:2546.Google Scholar
Ramirez, JM, Doi, A, Garcia, AJ, Elsen, FP, Koch, H, Wei, AD. The cellular building blocks of breathing. Compr Physiol, 2012; 2:2683–731.CrossRefGoogle ScholarPubMed
Subramanian, HH, Holstege, G. Stimulation of the midbrain periaqueductal gray modulates preinspiratory neurons in the ventrolateral medulla in the rat in vivo. J Comp Neurol, 2013; 521:3083–98.Google Scholar
Cruz-Sanchez, FF, Lucena, J, Ascaso, C, Tolosa, E, Quinto, L, Rossi, ML. Cerebellar cortex delayed maturation in Sudden Infant Death Syndrome. J Neuropathol Exp Neurol, 1997; 56, 340–6.Google Scholar
Lavezzi, AM, Ottaviani, G, Mauri, M, Matturri, L. Alterations of biological features of the cerebellum in sudden perinatal and infant death. Curr Mol Med, 2006; 6:429–35.Google Scholar
Calton, MA, Howard, JR, Harper, RM, Goldowitz, D, Mittleman, G. The cerebellum and SIDS: disordered breathing in a mouse model of developmental cerebellar Purkinje cell loss during recovery from hypercarbia. Front Neurol, 2016; 7:78.Google Scholar
Kinney, HC, Cryan, JB, Haynes, RL, Paterson, DS, Haas, EA, Mena, OJ, et al. Dentate gyrus abnormalities in sudden unexplained death in infants: morphological marker of underlying brain vulnerability. Acta Neuropathol, 2015; 129:6580.CrossRefGoogle ScholarPubMed
Hefti, MM, Kinney, HC, Cryan, JB, Haas, EA, Chadwick, AE, Crandall, LA, et al. Sudden unexpected death in early childhood: general observations in a series of 151 cases: Part 1 of the investigations of the San Diego SUDC Research Project. Forensic Sci Med Pathol, 2016; 12:413.Google Scholar
Ramirez, JM. The human pre-Bötzinger complex identified. Brain, 2011; 134:810.Google Scholar
McKay, LC, Janczewski, WA, Feldman, JL. Sleep-disordered breathing after targeted ablation of preBötzinger complex neurons. Nat Neurosci, 2005; 8:1142–4.Google Scholar
Ramirez, JM, Schwarzacher, SW, Pierrefiche, O, Olivera, BM, Richter, DW. Selective lesioning of the cat pre-Bötzinger complex in vivo eliminates breathing but not gasping. J Physiol, 1998; 507(Pt 3):895907.Google Scholar
Wenninger, JM, Pan, LG, Klum, L, Leekley, T, Bastastic, J, Hodges, MR, et al. Large lesions in the pre-Bötzinger complex area eliminate eupneic respiratory rhythm in awake goats. J Appl Physiol (1985), 2004; 97:1629–36.Google Scholar
Naeye, RL, Ladis, B, Drage, JS. Sudden Infant Death Syndrome. A prospective study.Am J Dis Child, 1976; 130:1207–10.Google Scholar
Gray, PA, Rekling, JC, Bocchiaro, CM, Feldman, JL. Modulation of respiratory frequency by peptidergic input to rhythmogenic neurons in the preBötzinger complex. Science, 1999; 286:1566–8.CrossRefGoogle ScholarPubMed
Stornetta, RL, Rosin, DL, Wang, H, Sevigny, CP, Weston, MC, Guyenet, PG. A group of glutamatergic interneurons expressing high levels of both neurokinin-1 receptors and somatostatin identifies the region of the pre-Bötzinger complex. J Comp Neurol, 2003; 455:499512.Google Scholar
Bouvier, J, Thoby-Brisson, M, Renier, N, Dubreuil, V, Ericson, J, Champagnat, J, et al. Hindbrain interneurons and axon guidance signaling critical for breathing. Nat Neurosci, 2010; 13:1066–74Google Scholar
Gray, PA, Hayes, JA, Ling, GY, Llona, I, Tupal, S, Picardo, MC, et al. Developmental origin of preBötzinger complex respiratory neurons. J Neurosci, 2010; 30:14883–95.Google Scholar
Wang, X, Hayes, JA, Revill, AL, Song, H, Kottick, A, Vann, NC, et al. Laser ablation of Dbx1 neurons in the pre-Bötzinger complex stops inspiratory rhythm and impairs output in neonatal mice. eLife, 2014; 3:e03427.Google Scholar
Revill, AL, Vann, NC, Akins, VT, Kottick, A, Gray, PA, Del Negro, CA, Funk, GD. Dbx1 precursor cells are a source of inspiratory XII premotoneurons. eLife, 2015; 4:e12301; https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4764567/; (accessed 31 October 2018).Google Scholar
Koizumi, H, Mosher, B, Tariq, MF, Zhang, R, Koshiya, N, Smith, JC. Voltage-dependent rhythmogenic property of respiratory pre-Bötzinger complex glutamatergic, Dbx1-derived, and somatostatin-expressing neuron populations revealed by graded optogenetic inhibition. eNeuro, 2016; 3(3); https://doi.org/10.1523/ENEURO.0081-16.2016 (accessed 31 October 2018).CrossRefGoogle ScholarPubMed
Vann, NC, Pham, FD, Hayes, JA, Kottick, A, Del Negro, CA. Transient suppression of Dbx1 preBötzinger interneurons disrupts breathing in adult mice. PLoS One, 2016; 11:e0162418; https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5017730/ (accessed 31 October 2018).Google Scholar
Sherman, D, Worrell, JW, Cui, Y, Feldman, JL. Optogenetic perturbation of preBötzinger complex inhibitory neurons modulates respiratory pattern. Nat Neurosci, 2015; 18:408–14.Google Scholar
Winter, SM, Fresemann, J, Schnell, C, Oku, Y, Hirrlinger, J, Hulsmann, S. Glycinergic interneurons are functionally integrated into the inspiratory network of mouse medullary slices. Pflugers Arch, 2009; 458:459–69.Google Scholar
Koch, H, Caughie, C, Elsen, FP, Doi, A, Garcia, AJ, Zanella, S, Ramirez, JM. Prostaglandin E2 differentially modulates the central control of eupnoea, sighs, and gasping in mice. J Physiol, 2015; 593:305–19Google Scholar
Tryba, AK, Pena, F, Lieske, SP, Viemari, JC, Thoby-Brisson, M, Ramirez, JM. Differential modulation of neural network and pacemaker activity underlying eupnea and sigh-breathing activities. J Neurophysiol, 2008; 99:2114–25Google Scholar
Doi, A, Ramirez, JM. State-dependent interactions between excitatory neuromodulators in the neuronal control of breathing. J Neurosci, 2010; 30:8251–62.Google Scholar
Pena, F, Ramirez, JM. Substance P-mediated modulation of pacemaker properties in the mammalian respiratory network. J Neurosci, 2004; 24:7549–56.Google Scholar
Pena, F, Ramirez, JM. Endogenous activation of serotonin-2A receptors is required for respiratory rhythm generation in vitro. J Neurosci, 2002; 22:11055–64.Google Scholar
Erickson, JT, Shafer, G, Rossetti, MD, Wilson, CG, Deneris, ES Arrest of 5-HT neuron differentiation delays respiratory maturation and impairs neonatal homeostatic responses to environmental challenges. Respir Physiol Neurobiol. 2007; 159(1):85101.Google Scholar
Cummings, KJ, Hewitt, JC, Li, A, Daubenspeck, JA, Nattie, EE Postnatal loss of brainstem serotonin neurones compromises the ability of neonatal rats to survive episodic severe hypoxia. J Physiol. 2011; 589(21):5247–56.Google Scholar
Buchanan, GF, Richerson, GB Central serotonin neurons are required for arousal to CO2. Proc Natl Acad Sci USA, 2010; 107(37):16354–9.CrossRefGoogle ScholarPubMed
Hodges, MR, Wehner, M, Aungst, J, Smith, JC, Richerson, GB Transgenic mice lacking serotonin neurons have severe apnea and high mortality during development. J Neurosci. 2009; 29(33):10341–9.Google Scholar
Tryba, AK, Pena, F, Ramirez, JM. Gasping activity in vitro: a rhythm dependent on 5-HT2A receptors. J Neurosci, 2006; 26:2623–34.Google Scholar
Puissant, MM, Mouradian, GC, Liu, P, Hodges, MR Identifying candidate genes that underlie cellular pH sensitivity in serotonin neurons using transcriptomics: a potential role for Kir5.1 channels. Front Cell Neurosci, 2017; 11:34; https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5318415/ (accessed 31 October 2018).Google Scholar
Viemari, JC, Garcia, AJ, Doi, A, Elsen, G, Ramirez, JM. beta-Noradrenergic receptor activation specifically modulates the generation of sighs in vivo and in vitro. Front Neural Circuits, 2013; 7:179.Google Scholar
Viemari, JC, Garcia, AJ, Doi, A, Ramirez, JM. Activation of alpha-2 noradrenergic receptors is critical for the generation of fictive eupnea and fictive gasping inspiratory activities in mammals in vitro. Eur J Neurosci, 2011; 33:2228–37.Google Scholar
Li, A, Nattie, E. Antagonism of rat orexin receptors by almorexant attenuates central chemoreception in wakefulness in the active period of the diurnal cycle. J Physiol, 2010; 588:2935–44.Google Scholar
Li, P, Janczewski, WA, Yackle, K, Kam, K, Pagliardini, S, Krasnow, MA, Feldman, JL. The peptidergic control circuit for sighing. Nature, 2016; 530:293–7.Google Scholar
Levy, O. Innate immunity of the newborn: basic mechanisms and clinical correlates. Nat Rev Immunol, 2007; 7(5):379–90.Google Scholar
Walker, JC, Smolders, MAJC, Gemen, EFA, Antonius, TAJ, Leuvenink, J, de Vries, E. Development of lymphocyte subpopulations in preterm infants. Scand J Immunol, 2011; 73(1):53–8.Google Scholar
An, G, Nieman, G, Vodovotz, Y. Toward computational identification of multiscale ‘tipping points’ in acute inflammation and multiple organ failure. Ann Biomed Eng, 2012; 40(11):2414–24.Google Scholar
Mi, Q, Constantine, G, Ziraldo, C, Solovyev, A, Torres, A, Namas, R, Bentley, T, et al. A dynamic view of trauma/hemorrhage-induced inflammation in mice: principal drivers and networks. PLoS One, 2011; 6(5):e19424.Google Scholar
Balan, KV, Kc, P, Hoxha, Z, Mayer, CA, Wilson, CG, Martin, RJ Vagal afferents modulate cytokine-mediated respiratory control at the neonatal medulla oblongata. Respir Physiol Neurobiol, 2011; 178(3):458–64.Google Scholar
Gresham, K, Boyer, B, Mayer, CA, Foglyano, R, Martin, R, Wilson, CG Airway inflammation and central respiratory control: results from in vivo and in vitro neonatal rat. Respir Physiol Neurobiol, 2011; 178(3):414–21.Google Scholar
Huxtable, AG, Vinit, S, Windelborn, JA, Crader, SM, Guenther, CG, Watters, JJ, Mitchell, GS Systemic inflammation impairs respiratory chemoreflexes and plasticity. Respir Physiol Neurobiol, 2011; 178(3):482–9.Google Scholar
Vinit, S, Windelborn, JA, Mitchell, GS Lipopolysaccharide attenuates phrenic long-term facilitation following acute intermittent hypoxia. Respir Physiol Neurobiol, 2011; 176(3):130–5.Google Scholar
Olsson, A, Kayhan, G, Lagercrantz, H, Herlenius, E. IL-1 beta depresses respiration and anoxic survival via a prostaglandin-dependent pathway in neonatal rats. Pediatr Res, 2003; 54(3):326–31.Google Scholar
Hofstetter, AO, Saha, S, Siljehav, V, Jakobsson, P, Herlenius, E. The induced prostaglandin e2 pathway is a key regulator of the respiratory response to infection and hypoxia in neonates. Proc Natl Acad Sci USA, 2007; 104(23):9894–9.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.

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.

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.

Available formats
×