Comprehensive Physiology Wiley Online Library

Central Chemoreceptors: Locations and Functions

Full Article on Wiley Online Library



Abstract

Central chemoreception traditionally refers to a change in ventilation attributable to changes in CO2/H+ detected within the brain. Interest in central chemoreception has grown substantially since the previous Handbook of Physiology published in 1986. Initially, central chemoreception was localized to areas on the ventral medullary surface, a hypothesis complemented by the recent identification of neurons with specific phenotypes near one of these areas as putative chemoreceptor cells. However, there is substantial evidence that many sites participate in central chemoreception some located at a distance from the ventral medulla. Functionally, central chemoreception, via the sensing of brain interstitial fluid H+, serves to detect and integrate information on (i) alveolar ventilation (arterial PCO2), (ii) brain blood flow and metabolism, and (iii) acid‐base balance, and, in response, can affect breathing, airway resistance, blood pressure (sympathetic tone), and arousal. In addition, central chemoreception provides a tonic “drive” (source of excitation) at the normal, baseline PCO2 level that maintains a degree of functional connectivity among brainstem respiratory neurons necessary to produce eupneic breathing. Central chemoreception responds to small variations in PCO2 to regulate normal gas exchange and to large changes in PCO2 to minimize acid‐base changes. Central chemoreceptor sites vary in function with sex and with development. From an evolutionary perspective, central chemoreception grew out of the demands posed by air versus water breathing, homeothermy, sleep, optimization of the work of breathing with the “ideal” arterial PCO2, and the maintenance of the appropriate pH at 37°C for optimal protein structure and function. © 2012 American Physiological Society. Compr Physiol 2:221‐254, 2012.

Comprehensive Physiology offers downloadable PowerPoint presentations of figures for non-profit, educational use, provided the content is not modified and full credit is given to the author and publication.

Download a PowerPoint presentation of all images


Figure 1. Figure 1.

The response of alveolar ventilation to CO2 inhalation in normal acid‐base conditions (X) as well as in chronic metabolic acidosis (solid triangles) and alkalosis (solid circles) in a conscious goat. Fencl et al., Am. J. Physiol. 1966 65, used with permission.

Figure 2. Figure 2.

Effect of light anesthesia on the ventilatory response to inhaled 2.56% and 5.6% CO2. While breathing 5.6% CO2, 10 mg/kg sodium pentothal was injected through a catheter in the jugular vein. Subsequent measurements were made during the last 5 min of a 15‐min period of anesthesia. Pappenheimer et al., Am. J. Physiol. 1965 201, used with permission.

Figure 3. Figure 3.

Locations of central chemoreceptors; the classic view: chemoreception located at ventral medullary surface (left panel) and the current view: chemoreception is widely distributed in hindbrain (right panel). Abbreviations: R, rostral; M, middle; C, caudal; LHA, lateral hypothalamus; DR, dorsal raphe; FN, fastigial nucleus; 4v, fourth ventricle; LC, locus ceruleus; 7N, facial nerve; cNTS, caudal nucleus tractus solitarious; AMB, ambiguous; VII, facial nucleus; SO, superior olive; PBC, pre‐Bötzinger Complex; rVRG, rostral ventral respiratory group; cVLM, caudal ventrolateral medulla; RTH/pFRG, retrotrapezoid nucleus/parafacial respiratory group; and Pn, pons [modified from Figure 1 in Nattie 168 and used with permission].

Figure 4. Figure 4.

A confocal image of arteries filled with fluorescein‐tagged albumin (red) and serotonergic neurons stained with anti‐TPH antibody (green). (A) TPH‐IR neurons and arteries seen in on the ventral surface of the medulla en bloc; filled vessels include arteries and some veins. [Adapted with permission from Macmillan Publishers Ltd: Nature Neuroscience 20, 2002.]

Figure 5. Figure 5.

Drawing of experimental setup including blow up of dialysis probe tip (A) and example of typical tracings (breathing, electroencephalogram (EEG), and neck electromyogram (EMG)) in air and 7% CO2 during wakefulness (AW), nonrapid eye movement (NREM), and rapid eye movement (REM) sleep (B).

Figure 6. Figure 6.

The average response of integrated phrenic amplitude to increased end‐tidal PCO2 before (solid line) and after (dotted line) injection of 100 nl kainic acid into the retrotrapezoid nucleus (RTN). These rats were anesthetized initially with halothane followed by chloralose‐urethane. Mean +/− SEM values are shown (n = 4).

Reprinted from Respiration Physiology, 97, Nattie EE, and Li A. RTN lesions decrease phrenic activity and CO2 sensitivity in rats. 63‐77, 1994, 182 with permission from Elsevier.
Figure 7. Figure 7.

Ventilation in absolute terms (A) and expressed as % baseline (B) in unanesthetized rats (n = 7) dialyzed with 25% CO2 in the retrotrapezoid nucleus during wakefulness (solid circles n = 13 trials) and behaviorally defined sleep (open circles, n = 10 trials). Mean +/− SEM values are shown. Control room air values were obtained before and after 20‐min period of dialysis. The four preexposure control values were combined into single value. Ventilation during focal RTN acidification was significantly greater during wakefulness when expressed in absolute terms or as % baseline. Note that ventilation increased to 24% of baseline. There was no response during sleep. Li et al., J. Appl. Physiol. 1999, 142, used with permission.

Figure 8. Figure 8.

Location and general characteristics of retrotrapezoid nucleus (RTN) neurons. (A) Schematic but correctly scaled drawing of a parasagittal section through the pontomedullary region of the adult rat showing the location of the RTN. BötC, Bötzinger region of the ventral respiratory column; PBC, pre‐Bötzinger region; rVRG, rostral ventral respiratory group; cVRG, caudal ventral respiratory group; IS, inferior salivary nucleus; LRt, lateral reticular nucleus; nA, nucleus ambiguus pars compacta; SO, superior olive; tz, trapezoid body; 7, facial motor nucleus; and 7n, seventh nerve. The two black dots are the cell bodies of the neurons shown in E and F. (B) Coronal section at the level indicated by the arrows in A showing the distribution of neurons that express Phox2b (left side of brain). The chemoreceptors are the Phox2b‐positive neurons that do not express tyrosine‐hydroxylase (TH). The cells that express both markers are the C1 neurons, which regulate blood pressure. C Method used to record from RTN neurons in vivo. D Effect of changing end‐expiratory CO2 on the activity of an RTN neuron recorded in vivo after intracerebral injection of the glutamate blocker kynurenic acid. The neuron still encodes the level of arterial CO2 despite the fact that the drug has silenced the activity of the central pattern generator (CPG; evidence that it has is not shown in the figure). (E and F) Structure of two RTN neurons recorded in vivo illustrating the fact that a major portion of the dendritic domain of these cells resides within the marginal layer of the ventral medullary surface. Guyenet, PG, J. Appl. Physiol. 2008, 82, used with permission.

Figure 9. Figure 9.

Effect of bilateral injections of SSP‐SAP into the RTN on the central chemoreflex. (A) Relationship between phrenic nerve discharge (PND) and end‐expiratory CO2 in a control rat two weeks after bilateral injection of saline. The apnoeic threshold is 5.2%. (B) Same experiment in a different rat two weeks after bilateral treatment with 2 × 0.6 ng of SSP‐SAP. The apnoeic threshold is 7.9%. PND above the apnoeic threshold appears normal. (C) Relationship between mv PND and end‐expiratory CO2 in controls (n = 19) and in 11 rats treated bilaterally with 2 × 0.6 ng of SSP‐SAP causing the destruction of 70% of the Phox2b+TH neurons of RTN. One arbitrary unit represents the highest value of mv PND registered at steady state with end‐expiratory CO2 set at 9.5% to 10%. *Statistical significance by RM ANOVA (P < 0.05). (D) Effect of graded lesions of the Phox2b+TH neurons of the RTN on the apnoeic threshold measured as shown in A and B. *Statistically significant difference from the other two groups by ANOVA (P < 0.05). (E) Correlation between apnoeic threshold and percentage Phox2b+TH neurons remaining (10 rats with two injections of toxin on each side and six rats with one injection on each side). The F, r2, and probability values of the linear regression are indicated in the figure. Takakura et al., J. Physiol., John Wiley and Sons 253, used with permission.

Figure 10. Figure 10.

Neurons in cell cultures from the medullary raphe are chemosensitive to acidosis. (A) Example of the firing rate of an acidosis‐stimulated neuron in response to respiratory acidosis and alkalosis. Lower trace is bath pH measured simultaneously at the inflow to the recording chamber. (B) Example of the firing rate of an acidosis‐inhibited neuron in response to the same stimuli. (C) Acidosis‐stimulated neurons respond to both respiratory acidosis and metabolic acidosis, indicating that a change in pHo (and/or intracellular pH), in the absence of changes in CO2, is sufficient for a response to occur.

Reprinted from Respiration Physiology, 129, Richerson et al., Chemosensitivity of serotonergic neurons of the rostral ventral medulla. 175‐189, 2001, 219, with permission from Elsevier.
Figure 11. Figure 11.

Focal acidification just below the caudal ventral surface chemosensitive area increases ventilation. Illustrative example of a single experiment in a responsive animal. The period of focal acidification is depicted by the gray rectangle with control periods shown before and after. The EEG and EMG recordings shown in A indicate the presence of wakefulness and sleep periods before, during and after focal acidosis. The ventilation data in the bottom of A shows that the increase of ventilation during high CO2 dialysis is present only in wakefulness. Note that at the onset of high CO2 dialysis when the rat was asleep (open symbols) ventilation did not increase but quickly did so when the animal woke up (filled symbols). In B, we show actual recordings of the plethysmograph pressure signal (upper trace), the EEG (middle trace), and the EMG (lower trace) taken from the averaged data points depicted by the lines and arrows.

Reprinted from Respiration Physiology & Neurobiology, 171, da Silva et al., High CO2/H+ dialysis in the caudal ventrolateral medulla (Loeschcke's area) increases ventilation in wakefulness. 46‐53, 2010, 36, with permission from Elsevier.
Figure 12. Figure 12.

The role of orexin in central chemoreception. The top panel is a schematic of a saggital section of rat brain showing the location of orexin containing neurons in the hypothalamus and their projection sites in the retrotrapezoid nucleus (RTN) and raphe magnus (RM) and raphe obscurus (Rob), which have been identified as participating in central chemoreception. The arrows point to plots demonstrating chemoreception effects at each site, At the left, the hypercapnic responses of ventilation in wild‐type (WT) mice and prepro‐orexin knockout mice (ORX‐KO) during quiet Wakefulness are shown. Data are presented as means ± SEM of five WT mice and five ORX‐KO mice. *P < 0.05 compared with WT mice. At the right, the Figure shows the effects of dialysis of vehicle solution (solid circles; N = 6) and 5 mM SB‐334867, an Ox1R antagonist, (open circles; N = 6) into the medullary raphe (MR) on ventilation while the rats were breathing air and 7% CO2 during wakefulness in the dark period. Mean ± SEM values are shown. The −16% effect is significant comparing vehicle to SB‐334867 treatment during 7% CO2 breathing for ventilation (P < 0.001, repeated measures ANOVA interaction with gas type). In the middle, the Figure shows the effects of dialysis of vehicle solution (filled circles; n = 6) and 5 mM SB‐334867 (open circles; n = 6) into the RTN on ventilation while the rats were breathing air and 7% CO2 during wakefulness. Mean ± SEM values are shown. The −30% effect is significant comparing vehicle to SB‐334867 treatment during 7% CO2 breathing (P < 0.01 post hoc comparison). This composite is modified from Figure 1 in Nattie 168, and used with permission (top), Figure 3 from Respiration Physiology & Neurobiology, 164, Kuwacki, Orexinergic modulation of breathing across vigilance states, 204‐212, 2008, 129 and used with permission of Elsevier (bottom left), Figure 2 from Respiration Physiology & Neurobiology, 170, Dias et al., The orexin receptor 1 (OX1R) in the rostral MR contributes to the hypercapnic chemoreflex in wakefulness, during the active period of the diurnal cycle, 96‐102, 2010, 48 and used with permission of Elsevier (bottom right), and Figure 2 in Dias et al., 49 J. Physiol. used with permission, John Wiley and Sons (bottom middle).

Figure 13. Figure 13.

The relationship between pulmonary ventilation (VE) and the arterial partial pressure of CO2 (PaCO2) of sham and rats treated with 6‐OHDA‐induced lesions (>50%) of noradrenergic neurons of the locus ceruleus exposed to normocapnia (0% CO2) and hypercapnia (7% CO2). The overall effect was a 64% reduction in the CO2 response. With kind permission from Springer Science and Business Media: Pflügers Archiv, Locus coeruleus noradrenergic neurons and CO2 drive to breathing, 455, 2008, pp. 1119‐28, Biancardi, V., Bicego, K. C., Almeida, M. C., and Gargaglioni, L. H Figure 3. 18.

Figure 14. Figure 14.

The arterial partial pressure of CO2 (PaCO2) in mm Hg is shown as a function of alveolar ventilation relative to a “normal” value (that associated with PaCO2 = 40 mm Hg) of 1.0 and multiples thereof. A constant and normal metabolic rate is assumed. The inset magnifies the central region of the plot. As alveolar ventilation increases, PaCO2 decreases, and vice versa. A 10% change in alveolar ventilation is associated with a 4 mm Hg change in PaCO2.

Figure 15. Figure 15.

Responses of VE, VT, and f to 0% to 7.6% CO2 in eight awake, unrestrained rats. Note that the response at low inspired CO2 levels is substantial such that PaCO2 is unchanged, that is, PaCO2 regulation is “perfect”. Lai et al., J. Appl. Physiol. 1978, 132, used with permission.

Figure 16. Figure 16.

Ventilatory response, expressed as % baseline, to focal acidification of: (A) the retrotrapezoid nucleus (RTN) alone [with simultaneous acidification of sites lying outside the raphe obscurus (ROb)], (B) the ROb alone (with simultaneous acidification of the sites lying outside the RTN), and (A and B) both the RTN and the ROb simultaneously. Asterisk at far right of A and of A and B indicates a significant (P < 0.01) increase as determined by repeated‐measures ANOVA on the absolute values for ventilation. Asterisks at individual time periods indicate a significant (P < 0.05) difference from baseline as determined by post hoc comparison with Dunnett's test. Modified from Figures 3, 4, and 5 in Dias et al., J. Appl. Physiol. 2008, 47, and used with permission.

Figure 17. Figure 17.

A schematic model for central chemoreception in wakefulness that represents our current working hypothesis. The areas in red represent sites at which focal acidification by dialysis with aCSF equilibrated with high CO2 produced an increase in ventilation in wakefulness. The areas in gray represent sites at which we anticipate a response to focal acidification in wakefulness. Solid lines show established functional connections related to chemoreception, for example, dialysis of an OX1R antagonist at the RTN decreased the CO2 response in wakefulness. Dotted lines show likely connections that remain to be established. That linking the caudal MR to the PBC reflects observations obtained in a slice preparation. The red line between the caudal MR and the RTN reflects a CO2 linked connection, that is, focal acidification of the caudal MR enhances the response to focal acidification of the RTN. Abbreviations are: LH, a designation that includes orexin neurons in lateral hypothalamus, dorsomedial hypothalamus and perifornical area; LC, locus ceruleus; CB, carotid body; NTS, nucleus tractus solitarius; PBC, pre‐Bötzinger complex; RTN, retrotrapezoid nucleus; MR, medullary raphe; and VLM, ventrolateral medulla. Nattie & Li, 175 J. Appl. Physiol. 2010, used with permission.

Figure 18. Figure 18.

A schematic model for central chemoreception in NREM (nonrapid eye movement) sleep that represents our current working hypothesis. The symbols and lines are as in Figure 17.Nattie & Li, 175 J. Appl. Physiol. 2010, used with permission.



Figure 1.

The response of alveolar ventilation to CO2 inhalation in normal acid‐base conditions (X) as well as in chronic metabolic acidosis (solid triangles) and alkalosis (solid circles) in a conscious goat. Fencl et al., Am. J. Physiol. 1966 65, used with permission.



Figure 2.

Effect of light anesthesia on the ventilatory response to inhaled 2.56% and 5.6% CO2. While breathing 5.6% CO2, 10 mg/kg sodium pentothal was injected through a catheter in the jugular vein. Subsequent measurements were made during the last 5 min of a 15‐min period of anesthesia. Pappenheimer et al., Am. J. Physiol. 1965 201, used with permission.



Figure 3.

Locations of central chemoreceptors; the classic view: chemoreception located at ventral medullary surface (left panel) and the current view: chemoreception is widely distributed in hindbrain (right panel). Abbreviations: R, rostral; M, middle; C, caudal; LHA, lateral hypothalamus; DR, dorsal raphe; FN, fastigial nucleus; 4v, fourth ventricle; LC, locus ceruleus; 7N, facial nerve; cNTS, caudal nucleus tractus solitarious; AMB, ambiguous; VII, facial nucleus; SO, superior olive; PBC, pre‐Bötzinger Complex; rVRG, rostral ventral respiratory group; cVLM, caudal ventrolateral medulla; RTH/pFRG, retrotrapezoid nucleus/parafacial respiratory group; and Pn, pons [modified from Figure 1 in Nattie 168 and used with permission].



Figure 4.

A confocal image of arteries filled with fluorescein‐tagged albumin (red) and serotonergic neurons stained with anti‐TPH antibody (green). (A) TPH‐IR neurons and arteries seen in on the ventral surface of the medulla en bloc; filled vessels include arteries and some veins. [Adapted with permission from Macmillan Publishers Ltd: Nature Neuroscience 20, 2002.]



Figure 5.

Drawing of experimental setup including blow up of dialysis probe tip (A) and example of typical tracings (breathing, electroencephalogram (EEG), and neck electromyogram (EMG)) in air and 7% CO2 during wakefulness (AW), nonrapid eye movement (NREM), and rapid eye movement (REM) sleep (B).



Figure 6.

The average response of integrated phrenic amplitude to increased end‐tidal PCO2 before (solid line) and after (dotted line) injection of 100 nl kainic acid into the retrotrapezoid nucleus (RTN). These rats were anesthetized initially with halothane followed by chloralose‐urethane. Mean +/− SEM values are shown (n = 4).

Reprinted from Respiration Physiology, 97, Nattie EE, and Li A. RTN lesions decrease phrenic activity and CO2 sensitivity in rats. 63‐77, 1994, 182 with permission from Elsevier.


Figure 7.

Ventilation in absolute terms (A) and expressed as % baseline (B) in unanesthetized rats (n = 7) dialyzed with 25% CO2 in the retrotrapezoid nucleus during wakefulness (solid circles n = 13 trials) and behaviorally defined sleep (open circles, n = 10 trials). Mean +/− SEM values are shown. Control room air values were obtained before and after 20‐min period of dialysis. The four preexposure control values were combined into single value. Ventilation during focal RTN acidification was significantly greater during wakefulness when expressed in absolute terms or as % baseline. Note that ventilation increased to 24% of baseline. There was no response during sleep. Li et al., J. Appl. Physiol. 1999, 142, used with permission.



Figure 8.

Location and general characteristics of retrotrapezoid nucleus (RTN) neurons. (A) Schematic but correctly scaled drawing of a parasagittal section through the pontomedullary region of the adult rat showing the location of the RTN. BötC, Bötzinger region of the ventral respiratory column; PBC, pre‐Bötzinger region; rVRG, rostral ventral respiratory group; cVRG, caudal ventral respiratory group; IS, inferior salivary nucleus; LRt, lateral reticular nucleus; nA, nucleus ambiguus pars compacta; SO, superior olive; tz, trapezoid body; 7, facial motor nucleus; and 7n, seventh nerve. The two black dots are the cell bodies of the neurons shown in E and F. (B) Coronal section at the level indicated by the arrows in A showing the distribution of neurons that express Phox2b (left side of brain). The chemoreceptors are the Phox2b‐positive neurons that do not express tyrosine‐hydroxylase (TH). The cells that express both markers are the C1 neurons, which regulate blood pressure. C Method used to record from RTN neurons in vivo. D Effect of changing end‐expiratory CO2 on the activity of an RTN neuron recorded in vivo after intracerebral injection of the glutamate blocker kynurenic acid. The neuron still encodes the level of arterial CO2 despite the fact that the drug has silenced the activity of the central pattern generator (CPG; evidence that it has is not shown in the figure). (E and F) Structure of two RTN neurons recorded in vivo illustrating the fact that a major portion of the dendritic domain of these cells resides within the marginal layer of the ventral medullary surface. Guyenet, PG, J. Appl. Physiol. 2008, 82, used with permission.



Figure 9.

Effect of bilateral injections of SSP‐SAP into the RTN on the central chemoreflex. (A) Relationship between phrenic nerve discharge (PND) and end‐expiratory CO2 in a control rat two weeks after bilateral injection of saline. The apnoeic threshold is 5.2%. (B) Same experiment in a different rat two weeks after bilateral treatment with 2 × 0.6 ng of SSP‐SAP. The apnoeic threshold is 7.9%. PND above the apnoeic threshold appears normal. (C) Relationship between mv PND and end‐expiratory CO2 in controls (n = 19) and in 11 rats treated bilaterally with 2 × 0.6 ng of SSP‐SAP causing the destruction of 70% of the Phox2b+TH neurons of RTN. One arbitrary unit represents the highest value of mv PND registered at steady state with end‐expiratory CO2 set at 9.5% to 10%. *Statistical significance by RM ANOVA (P < 0.05). (D) Effect of graded lesions of the Phox2b+TH neurons of the RTN on the apnoeic threshold measured as shown in A and B. *Statistically significant difference from the other two groups by ANOVA (P < 0.05). (E) Correlation between apnoeic threshold and percentage Phox2b+TH neurons remaining (10 rats with two injections of toxin on each side and six rats with one injection on each side). The F, r2, and probability values of the linear regression are indicated in the figure. Takakura et al., J. Physiol., John Wiley and Sons 253, used with permission.



Figure 10.

Neurons in cell cultures from the medullary raphe are chemosensitive to acidosis. (A) Example of the firing rate of an acidosis‐stimulated neuron in response to respiratory acidosis and alkalosis. Lower trace is bath pH measured simultaneously at the inflow to the recording chamber. (B) Example of the firing rate of an acidosis‐inhibited neuron in response to the same stimuli. (C) Acidosis‐stimulated neurons respond to both respiratory acidosis and metabolic acidosis, indicating that a change in pHo (and/or intracellular pH), in the absence of changes in CO2, is sufficient for a response to occur.

Reprinted from Respiration Physiology, 129, Richerson et al., Chemosensitivity of serotonergic neurons of the rostral ventral medulla. 175‐189, 2001, 219, with permission from Elsevier.


Figure 11.

Focal acidification just below the caudal ventral surface chemosensitive area increases ventilation. Illustrative example of a single experiment in a responsive animal. The period of focal acidification is depicted by the gray rectangle with control periods shown before and after. The EEG and EMG recordings shown in A indicate the presence of wakefulness and sleep periods before, during and after focal acidosis. The ventilation data in the bottom of A shows that the increase of ventilation during high CO2 dialysis is present only in wakefulness. Note that at the onset of high CO2 dialysis when the rat was asleep (open symbols) ventilation did not increase but quickly did so when the animal woke up (filled symbols). In B, we show actual recordings of the plethysmograph pressure signal (upper trace), the EEG (middle trace), and the EMG (lower trace) taken from the averaged data points depicted by the lines and arrows.

Reprinted from Respiration Physiology & Neurobiology, 171, da Silva et al., High CO2/H+ dialysis in the caudal ventrolateral medulla (Loeschcke's area) increases ventilation in wakefulness. 46‐53, 2010, 36, with permission from Elsevier.


Figure 12.

The role of orexin in central chemoreception. The top panel is a schematic of a saggital section of rat brain showing the location of orexin containing neurons in the hypothalamus and their projection sites in the retrotrapezoid nucleus (RTN) and raphe magnus (RM) and raphe obscurus (Rob), which have been identified as participating in central chemoreception. The arrows point to plots demonstrating chemoreception effects at each site, At the left, the hypercapnic responses of ventilation in wild‐type (WT) mice and prepro‐orexin knockout mice (ORX‐KO) during quiet Wakefulness are shown. Data are presented as means ± SEM of five WT mice and five ORX‐KO mice. *P < 0.05 compared with WT mice. At the right, the Figure shows the effects of dialysis of vehicle solution (solid circles; N = 6) and 5 mM SB‐334867, an Ox1R antagonist, (open circles; N = 6) into the medullary raphe (MR) on ventilation while the rats were breathing air and 7% CO2 during wakefulness in the dark period. Mean ± SEM values are shown. The −16% effect is significant comparing vehicle to SB‐334867 treatment during 7% CO2 breathing for ventilation (P < 0.001, repeated measures ANOVA interaction with gas type). In the middle, the Figure shows the effects of dialysis of vehicle solution (filled circles; n = 6) and 5 mM SB‐334867 (open circles; n = 6) into the RTN on ventilation while the rats were breathing air and 7% CO2 during wakefulness. Mean ± SEM values are shown. The −30% effect is significant comparing vehicle to SB‐334867 treatment during 7% CO2 breathing (P < 0.01 post hoc comparison). This composite is modified from Figure 1 in Nattie 168, and used with permission (top), Figure 3 from Respiration Physiology & Neurobiology, 164, Kuwacki, Orexinergic modulation of breathing across vigilance states, 204‐212, 2008, 129 and used with permission of Elsevier (bottom left), Figure 2 from Respiration Physiology & Neurobiology, 170, Dias et al., The orexin receptor 1 (OX1R) in the rostral MR contributes to the hypercapnic chemoreflex in wakefulness, during the active period of the diurnal cycle, 96‐102, 2010, 48 and used with permission of Elsevier (bottom right), and Figure 2 in Dias et al., 49 J. Physiol. used with permission, John Wiley and Sons (bottom middle).



Figure 13.

The relationship between pulmonary ventilation (VE) and the arterial partial pressure of CO2 (PaCO2) of sham and rats treated with 6‐OHDA‐induced lesions (>50%) of noradrenergic neurons of the locus ceruleus exposed to normocapnia (0% CO2) and hypercapnia (7% CO2). The overall effect was a 64% reduction in the CO2 response. With kind permission from Springer Science and Business Media: Pflügers Archiv, Locus coeruleus noradrenergic neurons and CO2 drive to breathing, 455, 2008, pp. 1119‐28, Biancardi, V., Bicego, K. C., Almeida, M. C., and Gargaglioni, L. H Figure 3. 18.



Figure 14.

The arterial partial pressure of CO2 (PaCO2) in mm Hg is shown as a function of alveolar ventilation relative to a “normal” value (that associated with PaCO2 = 40 mm Hg) of 1.0 and multiples thereof. A constant and normal metabolic rate is assumed. The inset magnifies the central region of the plot. As alveolar ventilation increases, PaCO2 decreases, and vice versa. A 10% change in alveolar ventilation is associated with a 4 mm Hg change in PaCO2.



Figure 15.

Responses of VE, VT, and f to 0% to 7.6% CO2 in eight awake, unrestrained rats. Note that the response at low inspired CO2 levels is substantial such that PaCO2 is unchanged, that is, PaCO2 regulation is “perfect”. Lai et al., J. Appl. Physiol. 1978, 132, used with permission.



Figure 16.

Ventilatory response, expressed as % baseline, to focal acidification of: (A) the retrotrapezoid nucleus (RTN) alone [with simultaneous acidification of sites lying outside the raphe obscurus (ROb)], (B) the ROb alone (with simultaneous acidification of the sites lying outside the RTN), and (A and B) both the RTN and the ROb simultaneously. Asterisk at far right of A and of A and B indicates a significant (P < 0.01) increase as determined by repeated‐measures ANOVA on the absolute values for ventilation. Asterisks at individual time periods indicate a significant (P < 0.05) difference from baseline as determined by post hoc comparison with Dunnett's test. Modified from Figures 3, 4, and 5 in Dias et al., J. Appl. Physiol. 2008, 47, and used with permission.



Figure 17.

A schematic model for central chemoreception in wakefulness that represents our current working hypothesis. The areas in red represent sites at which focal acidification by dialysis with aCSF equilibrated with high CO2 produced an increase in ventilation in wakefulness. The areas in gray represent sites at which we anticipate a response to focal acidification in wakefulness. Solid lines show established functional connections related to chemoreception, for example, dialysis of an OX1R antagonist at the RTN decreased the CO2 response in wakefulness. Dotted lines show likely connections that remain to be established. That linking the caudal MR to the PBC reflects observations obtained in a slice preparation. The red line between the caudal MR and the RTN reflects a CO2 linked connection, that is, focal acidification of the caudal MR enhances the response to focal acidification of the RTN. Abbreviations are: LH, a designation that includes orexin neurons in lateral hypothalamus, dorsomedial hypothalamus and perifornical area; LC, locus ceruleus; CB, carotid body; NTS, nucleus tractus solitarius; PBC, pre‐Bötzinger complex; RTN, retrotrapezoid nucleus; MR, medullary raphe; and VLM, ventrolateral medulla. Nattie & Li, 175 J. Appl. Physiol. 2010, used with permission.



Figure 18.

A schematic model for central chemoreception in NREM (nonrapid eye movement) sleep that represents our current working hypothesis. The symbols and lines are as in Figure 17.Nattie & Li, 175 J. Appl. Physiol. 2010, used with permission.

References
 1. Abbott SB, Stornetta RL, Fortuna MG, Depuy SD, West GH, Harris TE, Guyenet PG. Photostimulation of retrotrapezoid nucleus phox2b‐expressing neurons in vivo produces long‐lasting activation of breathing in rats. J Neurosci 29: 5806‐5819, 2009.
 2. Adamantidis AR, Zhang F, Aravanis AM, Deisseroth K, de Lecea L. Neural substrates of awakening probed with optogenetic control of hypocretin neurons. Nature 450: 420‐424, 2007.
 3. Ainslie PN, Duffin J. Integration of cerebrovascular CO2 reactivity and chemoreflex control of breathing: Mechanisms of regulation, measurement, and interpretation. Am J Physiol 296: R1473‐R1495, 2009.
 4. Akilesh MR, Kamper M, Li A, Nattie EE. Effects of unilateral lesions of retrotrapezoid nucleus on breathing in awake rats. J Appl Physiol 82: 469‐479, 1997.
 5. Amiel J, Dubreuil V, Ramanantsoa N, Fortin G, Gallego J, Brunet JF, Goridis C. PHOX2B in respiratory control: Lessons from congenital central hypoventilation syndrome and its mouse models. Respir Physiol Neurobiol 168: 125‐132, 2009.
 6. Antunes VR, Brailoiu GC, Kwok EH, Scruggs P, Dun NJ. Orexins/hypocretins excite rat sympathetic preganglionic neurons in vivo and in vitro. Am J Physiol 281: R1801‐R1807, 2001.
 7. Arita H, Ichikawa K, Kuwana S, Kogo N. Possible locations of pH‐dependent central chemoreceptors: Intramedullary regions with acidic shift of extracellular fluid pH during hypercapnia. Brain Res 485: 285‐293, 1989.
 8. Aston‐Jones G, Rajkowski J, Cohen J. Locus coeruleus and regulation of behavioral flexibility and attention. Prog Brain Res 126: 165‐182, 2000.
 9. Ayas NT, Brown R, Shea SA. Hypercapnia can induce arousal from sleep in the absence of altered respiratory mechanoreception. AmJ Respir Crit Care Med 162: 1004‐1008, 2000.
 10. Badr MS, Toiber F, Skatrud JB, Dempsey J. Pharyngeal narrowing/occlusion during central sleep apnea. J Appl Physiol 78: 1806‐1815, 1995.
 11. Bainton CR, Kirkwood PA. The effect of carbon dioxide on the tonic and the rhythmic discharges of expiratory bulbospinal neurones. J Physiol 296: 291‐314, 1979.
 12. Batsel HL. Activity of bulbar respiratory neurons during passive hyperventilation. Exp Neurol 19: 357‐374, 1967.
 13. Berger AJ, Cooney KA. Ventilatory effects of kainic acid injection of the ventrolateral solitary nucleus. J Appl Physiol 52: 131‐140, 1982.
 14. Berkenbosch A, van Beek JH, Olievier CN, De Goede J, Quanjer PH. Central respiratory CO2 sensitivity at extreme hypocapnia. Respir Physiol 55: 95‐102, 1984.
 15. Bernard DG, Li A, Nattie EE. Evidence for central chemoreception in the midline raphe. J Appl Physiol 80: 108‐115, 1996.
 16. Berridge CW, Waterhouse BD. The locus coeruleus‐noradrenergic system: Modulation of behavioral state and state‐dependent cognitive processes. Brain Res Brain Res Rev 42: 33‐84, 2003.
 17. Berssenbrugge A, Dempsey J, Iber C, Skatrud J, Wilson P. Mechanisms of hypoxia‐induced periodic breathing during sleep in humans. J Physiol 343: 507‐524, 1983.
 18. Biancardi V, Bicego KC, Almeida MC, Gargaglioni LH. Locus coeruleus noradrenergic neurons and CO2 drive to breathing. Pflugers Arch 455: 1119‐1128, 2008.
 19. Blain GM, Smith CA, Henderson KS, Dempsey JA. Contribution of the carotid body chemoreceptors to eupneic ventilation in the intact, unanesthetized dog. J Appl Physiol 106: 1564‐1573, 2009.
 20. Bradley SR, Pieribone VA, Wang W, Severson CA, Jacobs RA, Richerson GB. Chemosensitive serotonergic neurons are closely associated with large medullary arteries. Nat Neurosci 5: 401‐402, 2002.
 21. Brisbare‐Roch C, Dingemanse J, Koberstein R, Hoever P, Aissaoui H, Flores S, Mueller C, Nayler O, van Gerven J, de Haas SL, Hess P, Qiu C, Buchmann S, Scherz M, Weller T, Fischli W, Clozel M, Jenck F. Promotion of sleep by targeting the orexin system in rats, dogs and humans. Nat Med 13: 150‐155, 2007.
 22. Brunet JF, Pattyn A. Phox2 genes ‐ from patterning to connectivity. Curr Opin Genet Dev 12: 435‐440, 2002.
 23. Buchanan GF, Richerson GB. Central serotonin neurons are required for arousal to CO2. Proc Natl Acad Sci U S A 107: 16354‐16359, 2010.
 24. Carter ME, Borg JS, de Lecea L. The brain hypocretins and their receptors: Mediators of allostatic arousal. Curr Opin Pharmacol 9: 39‐45, 2009.
 25. Carter ME, Yizhar O, Chikahisa S, Nguyen H, Adamantidis A, Nishino S, Deisseroth K, de Lecea L. Tuning arousal with optogenetic modulation of locus coeruleus neurons. Nat Neurosci 13: 1526‐1533, 2010.
 26. Chapman RW, Santiago TV, Edelman NH. Effects of graded reduction of brain blood flow on chemical control of breathing. J Appl Physiol 47: 1289‐1294, 1979.
 27. Chemelli RM, Willie JT, Sinton CM, Elmquist JK, Scammell T, Lee C, Richardson JA, Williams SC, Xiong Y, Kisanuki Y, Fitch TE, Nakazato M, Hammer RE, Saper CB, Yanagisawa M. Narcolepsy in orexin knockout mice: Molecular genetics of sleep regulation. Cell 98: 437‐451, 1999.
 28. Coates EL, Li A, Nattie EE. Widespread sites of brain stem ventilatory chemoreceptors. J Appl Physiol 75: 5‐14, 1993.
 29. Coates EL, Li AH, Nattie EE. Acetazolamide on the ventral medulla of the cat increases phrenic output and delays the ventilatory response to CO2. J Physiol 441: 433‐451, 1991.
 30. Cohen MI. Tonic chemoreceptor input as the background for respiratory rhythm. In: Trouth CO, Millis RM, Kiwull‐Schone HF, Schlafke ME, editors. Ventral Brainstem Mechanisms and Control of Respiration and Blood Pressure. New York: Marcel Dekker, 1995, p. 797‐799.
 31. Cohen MI, Piercey MF, Gootman PM, Wolotsky P. Respiratory rhythmicity in the cat. Fed Proc 35: 1967‐1974, 1976.
 32. Corcoran AE, Hodges MR, Wu Y, Wang W, Wylie CJ, Deneris ES, Richerson GB. Medullary serotonin neurons and central CO2 chemoreception. Respir Physiol Neurobiol 168: 49‐58, 2009.
 33. Cream C, Li A, Nattie E. The retrotrapezoid nucleus (RTN): Local cytoarchitecture and afferent connections. Respir Physiol Neurobiol 130: 121‐137, 2002.
 34. Cummings KJ, Commons KG, Fan KC, Li A, Nattie EE. Severe spontaneous bradycardia associated with respiratory disruptions in rat pups with fewer brain stem 5‐HT neurons. Am J Physiol 296: R1783‐R1796, 2009.
 35. Cummings KJ, Li A, Deneris ES, Nattie EE. Bradycardia in serotonin‐deficient Pet‐1‐/‐ mice: Influence of respiratory dysfunction and hyperthermia over the first 2 postnatal weeks. Am J Physiol Regul Integr Comp Physiol 298: R1333‐R1342, 2010.
 36. da Silva GS, Li A, Nattie E. High CO2/H+ dialysis in the caudal ventrolateral medulla (Loeschcke's area) increases ventilation in wakefulness. Respir Physiol Neurobiol 171: 46‐53, 2010.
 37. Dauger S, Pattyn A, Lofaso F, Gaultier C, Goridis C, Gallego J, Brunet JF. Phox2b controls the development of peripheral chemoreceptors and afferent visceral pathways. Development 130: 6635‐6642, 2003.
 38. Davis SE, Solhied G, Castillo M, Dwinell M, Brozoski D, Forster HV. Postnatal developmental changes in CO2 sensitivity in rats. J Appl Physiol 101: 1097‐1103, 2006.
 39. Dean JB. Theory of gastric CO2 ventilation and its control during respiratory acidosis: Implications for central chemosensitivity, pH regulation, and diseases causing chronic CO2 retention. Respir Physiol Neurobiol 175: 189‐209, 2011.
 40. Dean JB, Kinkade EA, Putnam RW. Cell‐cell coupling in CO(2)/H(+)‐excited neurons in brainstem slices. Respir Physiol 129: 83‐100, 2001.
 41. Dean JB, Lawing WL, Millhorn DE. CO2 decreases membrane conductance and depolarizes neurons in the nucleus tractus solitarii. Exp Brain Res 76: 656‐661, 1989.
 42. Dean JB, Nattie EE. Central CO2 chemoreception in cardiorespiratory control. J Appl Physiol 108: 976‐978, 2010.
 43. Dempsey JA, Forster HV. Mediation of ventilatory daptations. Physiol Rev 62: 262‐346, 1982.
 44. Dempsey JA, Smith CA, Przybylowski T, Chenuel B, Xie A, Nakayama H, Skatrud JB. The ventilatory responsiveness to CO(2) below eupnoea as a determinant of ventilatory stability in sleep. J Physiol 560: 1‐11, 2004.
 45. Deng BS, Nakamura A, Zhang W, Yanagisawa M, Fukuda Y, Kuwaki T. Contribution of orexin in hypercapnic chemoreflex: Evidence from genetic and pharmacological disruption and supplementation studies in mice. J Appl Physiol 103: 1772‐1779, 2007.
 46. Desarnaud F, Murillo‐Rodriguez E, Lin L, Xu M, Gerashchenko D, Shiromani SN, Nishino S, Mignot E, Shiromani PJ. The diurnal rhythm of hypocretin in young and old F344 rats. Sleep 27: 851‐856, 2004.
 47. Dias MB, Li A, Nattie E. Focal CO2 dialysis in raphe obscurus does not stimulate ventilation but enhances the response to focal CO2 dialysis in the retrotrapezoid nucleus. J Appl Physiol 105: 83‐90, 2008.
 48. Dias MB, Li A, Nattie E. The orexin receptor 1 (OX1R) in the rostral medullary raphe contributes to the hypercapnic chemoreflex in wakefulness, during the active period of the diurnal cycle. Respir Physiol Neurobiol 170: 96‐102, 2010.
 49. Dias MB, Li A, Nattie EE. Antagonism of orexin receptor‐1 in the retrotrapezoid nucleus inhibits the ventilatory response to hypercapnia predominantly in wakefulness. J Physiol 587: 2059‐2067, 2009.
 50. Dobbins EG, Feldman JL. Brainstem network controlling descending drive to phrenic motoneurons in rat. JComp Neurol 347: 64‐86, 1994.
 51. Dubreuil V, Barhanin J, Goridis C, Brunet JF. Breathing with phox2b. Philos Trans R Soc Lond B Biol Sci 364: 2477‐2483, 2009.
 52. Dubreuil V, Thoby‐Brisson M, Rallu M, Persson K, Pattyn A, Birchmeier C, Brunet JF, Fortin G, Goridis C. Defective respiratory rhythmogenesis and loss of central chemosensitivity in Phox2b mutants targeting retrotrapezoid nucleus neurons. J Neurosci 29: 14836‐14846, 2009.
 53. Duncan JR, Paterson DS, Hoffman JM, Mokler DJ, Borenstein NS, Belliveau RA, Krous HF, Haas EA, Stanley C, Nattie EE, Trachtenberg FL, Kinney HC. Brainstem serotonergic deficiency in sudden infant death syndrome. JAMA 303: 430‐437, 2010.
 54. Durand E, Dauger S, Pattyn A, Gaultier C, Goridis C, Gallego J. Sleep‐disordered breathing in newborn mice heterozygous for the transcription factor Phox2b. Am J Respir Crit Care Med 172: 238‐243, 2005.
 55. Dutschmann M, Kron M, Morschel M, Gestreau C. Activation of Orexin B receptors in the pontine Kolliker‐Fuse nucleus modulates pre‐inspiratory hypoglossal motor activity in rat. Respir Physiol Neurobiol 159: 232‐235, 2007.
 56. Elam M, Yao T, Thoren P, Svensson TH. Hypercapnia and hypoxia: Chemoreceptor‐mediated control of locus coeruleus neurons and splanchnic, sympathetic nerves. Brain Res 222: 373‐381, 1981.
 57. Eldridge FL, Kiley JP, Millhorn DE. Respiratory responses to medullary hydrogen ion changes in cats: Different effects of respiratory and metabolic acidoses. J Physiol 358: 285‐297, 1985.
 58. Elias CF, Saper CB, Maratos‐Flier E, Tritos NA, Lee C, Kelly J, Tatro JB, Hoffman GE, Ollmann MM, Barsh GS, Sakurai T, Yanagisawa M, Elmquist JK. Chemically defined projections linking the mediobasal hypothalamus and the lateral hypothalamic area. J Comp Neurol 402: 442‐459, 1998.
 59. Erlichman JS, Leiter JC. Glia modulation of the extracellular milieu as a factor in central CO2 chemosensitivity and respiratory control. J Appl Physiol 108: 1803‐1811, 2010.
 60. Erlichman JS, Leiter JC, Gourine AV. ATP, glia and central respiratory control. Respir Physiol Neurobiol 173: 305‐311, 2010.
 61. Erlichman JS, Li A, Nattie EE. Ventilatory effects of glial dysfunction in a rat brain stem chemoreceptor region. J Appl Physiol 85: 1599‐1604, 1998.
 62. Feldman JL, Gautier H. Interaction of pulmonary afferents and pneumotaxic center in control of respiratory pattern in cats. J Neurophysiol 39: 31‐44, 1976.
 63. Feldman JL, Mitchell GS, Nattie EE. Breathing: Rhythmicity, plasticity, chemosensitivity. Annu Rev Neurosci 26: 239‐266, 2003.
 64. Felten DL, Crutcher KA. Neuronal‐vascular relationships in the raphe nuclei, locus coeruleus, and substantia nigra in primates. Am J Anat 155: 467‐481, 1979.
 65. Fencl V, Miller TB, Pappenheimer JR. Studies on the respiratory response to disturbances of acid‐base balance, with deductions concerning the ionic composition of cerebral interstitial fluid. Am J Physiol 210: 459‐472, 1966.
 66. Fenik P, Veasey SC. Pharmacological characterization of serotonergic receptor activity in the hypoglossal nucleus. Am J Respir Crit Care Med 167: 563‐569, 2003.
 67. Fenik VB, Rukhadze I, Kubin L. Inhibition of pontine noradrenergic A7 cells reduces hypoglossal nerve activity in rats. Neuroscience 157: 473‐482, 2008.
 68. Fink BR. The stimulant effect of wakefulness on respiration: Clinical aspects. Br J Anaesth 33: 97‐101, 1961.
 69. Foote SL, Aston‐Jones G, Bloom FE. Impulse activity of locus coeruleus neurons in awake rats and monkeys is a function of sensory stimulation and arousal. Proc Natl Acad Sci U S A 77: 3033‐3037, 1980.
 70. Forster HV. Ventilatory effects of glial dysfunction in a rat brain stem chemoreceptor region. J Appl Physiol 85: 1597‐1598, 1998.
 71. Forster HV, Ohtake PJ, Pan LG, Lowry TF. Effect on breathing of surface ventrolateral medullary cooling in awake, anesthetized and asleep goats. Respir Physiol 110: 187‐197, 1997.
 72. Forster HV, Smith CA. Contributions of central and peripheral chemoreceptors to the ventilatory response to CO2/H+. J Appl Physiol 108: 989‐994, 2010.
 73. Fortuna MG, Stornetta RL, West GH, Guyenet PG. Activation of the retrotrapezoid nucleus by posterior hypothalamic stimulation. J Physiol 587: 5121‐5138, 2009.
 74. Fraigne JJ, Dunin‐Barkowski WL, Orem JM. Effect of hypercapnia on sleep and breathing in unanesthetized cats. Sleep 31: 1025‐1033, 2008.
 75. Fukuda Y, Honda Y, Schlafke ME, Loeschcke HH. Effect of H+ on the membrane potential of silent cells in the ventral and dorsal surface layers of the rat medulla in vitro. Pflugers Arch 376: 229‐235, 1978.
 76. Gahlenbeck H, Bartels H. Blood gas transport properties in gill and lung forms of the axolotl (ambystoma mexicanum). Respir Physiol 9: 175‐182, 1970.
 77. Gargaglioni LH, Hartzler LK, Putnam RW. The locus coeruleus and central chemosensitivity. Respir Physiol Neurobiol 173: 264‐273, 2010.
 78. Gestreau C, Heitzmann D, Thomas J, Dubreuil V, Bandulik S, Reichold M, Bendahhou S, Pierson P, Sterner C, Peyronnet‐Roux J, Benfriha C, Tegtmeier I, Ehnes H, Georgieff M, Lesage F, Brunet JF, Goridis C, Warth R, Barhanin J. Task2 potassium channels set central respiratory CO2 and O2 sensitivity. Proc Natl Acad Sci U S A 107: 2325‐2330.
 79. Gourine AV, Kasymov V, Marina N, Tang F, Figueiredo MF, Lane S, Teschemacher AG, Spyer KM, Deisseroth K, Kasparov S. Astrocytes control breathing through pH‐dependent release of ATP. Science 329: 571‐575, 2010.
 80. Gourine AV, Llaudet E, Dale N, Spyer KM. ATP is a mediator of chemosensory transduction in the central nervous system. Nature 436: 108‐111, 2005.
 81. Gourine AV, Wood JD, Burnstock G. Purinergic signalling in autonomic control. Trends Neurosci 32: 241‐248, 2009.
 82. Guyenet PG. The 2008 Carl Ludwig Lecture: Retrotrapezoid nucleus, CO2 homeostasis, and breathing automaticity. J Appl Physiol 105: 404‐416, 2008.
 83. Guyenet PG, Bayliss DA, Mulkey DK, Stornetta RL, Moreira TS, Takakura AT. The retrotrapezoid nucleus and central chemoreception. Adv Exp Med Biol 605: 327‐332, 2008.
 84. Guyenet PG, Bayliss DA, Stornetta RL, Fortuna MG, Abbott SB, DePuy SD. Retrotrapezoid nucleus, respiratory chemosensitivity and breathing automaticity. Respir Physiol Neurobiol 168: 59‐68, 2009.
 85. Guyenet PG, Mulkey DK, Stornetta RL, Bayliss DA. Regulation of ventral surface chemoreceptors by the central respiratory pattern generator. J Neurosci 25: 8938‐8947, 2005.
 86. Guyenet PG, Stornetta RL, Abbott SB, Depuy SD, Fortuna MG, Kanbar R. Central CO2 chemoreception and integrated neural mechanisms of cardiovascular and respiratory control. J Appl Physiol 108: 995‐1002, 2010.
 87. Guyenet PG, Stornetta RL, Bayliss DA. Central respiratory chemoreception. J Comp Neurol 518: 3883‐3906, 2010.
 88. Guyenet PG, Stornetta RL, Bayliss DA. Retrotrapezoid nucleus and central chemoreception. J Physiol 586: 2043‐2048, 2008.
 89. Hara J, Beuckmann CT, Nambu T, Willie JT, Chemelli RM, Sinton CM, Sugiyama F, Yagami K, Goto K, Yanagisawa M, Sakurai T. Genetic ablation of orexin neurons in mice results in narcolepsy, hypophagia, and obesity. Neuron 30: 345‐354, 2001.
 90. Hartzler LK, Dean JB, Putnam RW. The chemosensitive response of neurons from the locus coeruleus (LC) to hypercapnic acidosis with clamped intracellular pH. Adv Exp Med Biol 605: 333‐337, 2008.
 91. Haxhiu MA, Erokwu B, Bhardwaj V, Dreshaj IA. The role of the medullary raphe nuclei in regulation of cholinergic outflow to the airways. J Auton Nerv Syst 69: 64‐71, 1998.
 92. Haxhiu MA, Jansen AS, Cherniack NS, Loewy AD. CNS innervation of airway‐related parasympathetic preganglionic neurons: A transneuronal labeling study using pseudorabies virus. Brain Res 618: 115‐134, 1993.
 93. Haxhiu MA, Kc P, Neziri B, Yamamoto BK, Ferguson DG, Massari VJ. Catecholaminergic microcircuitry controlling the output of airway‐related vagal preganglionic neurons. J Appl Physiol 94: 1999‐2009, 2003.
 94. Haxhiu MA, Rust CF, Brooks C, Kc P. CNS determinants of sleep‐related worsening of airway functions: Implications for nocturnal asthma. Respir Physiol Neurobiol 151: 1‐30, 2006.
 95. Haxhiu MA, Yung K, Erokwu B, Cherniack NS. CO2‐induced c‐fos expression in the CNS catecholaminergic neurons. Respir Physiol 105: 35‐45, 1996.
 96. Hilaire G. Endogenous noradrenaline affects the maturation and function of the respiratory network: Possible implication for SIDS. Auton Neurosci 126‐127: 320‐331, 2006.
 97. Hilaire G, Viemari JC, Coulon P, Simonneau M, Bevengut M. Modulation of the respiratory rhythm generator by the pontine noradrenergic A5 and A6 groups in rodents. Respir Physiol Neurobiol 143: 187‐197, 2004.
 98. Hitzig BM, Perng WC, Burt T, Okunieff P, Johnson DC. 1H‐NMR measurement of fractional dissociation of imidazole in intact animals. Am J Physiol 266: R1008‐R1015, 1994.
 99. Hodges MR, Klum L, Leekley T, Brozoski DT, Bastasic J, Davis S, Wenninger JM, Feroah TR, Pan LG, Forster HV. Effects on breathing in awake and sleeping goats of focal acidosis in the medullary raphe. J Appl Physiol 96: 1815‐1824, 2004.
 100. Hodges MR, Martino P, Davis S, Opansky C, Pan LG, Forster HV. Effects on breathing of focal acidosis at multiple medullary raphe sites in awake goats. J Appl Physiol 97: 2303‐2309, 2004.
 101. Hodges MR, Richerson GB. Interaction between defects in ventilatory and thermoregulatory control in mice lacking 5‐HT neurons. Respir Physiol Neurobiol 164: 350‐357, 2008.
 102. Hodges MR, Tattersall GJ, Harris MB, McEvoy SD, Richerson DN, Deneris ES, Johnson RL, Chen ZF, Richerson GB. Defects in breathing and thermoregulation in mice with near‐complete absence of central serotonin neurons. J Neurosci 28: 2495‐2505, 2008.
 103. 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 29: 10341‐10349, 2009.
 104. Holleran J, Babbie M, Erlichman JS. Ventilatory effects of impaired glial function in a brain stem chemoreceptor region in the conscious rat. J Appl Physiol 90: 1539‐1547, 2001.
 105. Horner RL, Kozar LF, Kimoff RJ, Phillipson EA. Effects of sleep on the tonic drive to respiratory muscle and the threshold for rhythm generation in the dog. J Physiol 474: 525‐537, 1994.
 106. Horner RL, Kozar LF, Phillipson EA. Tonic respiratory drive in the absence of rhythm generation in the conscious dog. J Appl Physiol 76: 671‐680, 1994.
 107. Horner RL, Liu X, Gill H, Nolan P, Liu H, Sood S. Effects of sleep‐wake state on the genioglossus vs. diaphragm muscle response to CO(2) in rats. J Appl Physiol 92: 878‐887, 2002.
 108. Howell BJ, Baumgardner FW, Bondi K, Rahn H. Acid‐base balance in cold‐blooded vertebrates as a function of body temperature. Am J Physiol 218: 600‐606, 1970.
 109. Huckstepp RT, Id Bihi R, Eason R, Spyer KM, Dicke N, Willecke K, Marina N, Gourine AV, Dale N. Connexin hemichannel‐mediated CO2‐dependent release of ATP in the medulla oblongata contributes to central respiratory chemosensitivity. J Physiol 588: 3901‐3920, 2010.
 110. Hudgel DW, Hendricks C, Dadley A. Alteration in obstructive apnea pattern induced by changes in oxygen‐ and carbon‐dioxide‐inspired concentrations. Am Rev Respir Dis 138: 16‐19, 1988.
 111. Ichikawa K, Kuwana S, Arita H. ECF pH dynamics within the ventrolateral medulla: A microelectrode study. J Appl Physiol 67: 193‐198, 1989.
 112. Jacobs BL. Single unit activity of locus coeruleus neurons in behaving animals. Prog Neurobiol 27: 183‐194, 1986.
 113. Jelev A, Sood S, Liu H, Nolan P, Horner RL. Microdialysis perfusion of 5‐HT into hypoglossal motor nucleus differentially modulates genioglossus activity across natural sleep‐wake states in rats. J Physiol 532: 467‐481, 2001.
 114. Jensen P, Farago AF, Awatramani RB, Scott MM, Deneris ES, Dymecki SM. Redefining the serotonergic system by genetic lineage. Nat Neurosci 11: 417‐419, 2008.
 115. John J, Bailey EF, Fregosi RF. Respiratory‐related discharge of genioglossus muscle motor units. Am J Respir Crit Care Med 172: 1331‐1337, 2005.
 116. Johnson SM, Haxhiu MA, Richerson GB. GFP‐expressing locus ceruleus neurons from Prp57 transgenic mice exhibit CO2/H+ responses in primary cell culture. J Appl Physiol 105: 1301‐1311, 2008.
 117. Just JJ, Gatz RN, Crawford EC Jr. Changes in respiratory functions during metamorphosis of the bullfrog, Rana catesbeiana. Respir Physiol 17: 276‐282, 1973.
 118. Kanamaru M, Homma I. Compensatory airway dilation and additive ventilatory augmentation mediated by dorsomedial medullary 5‐hydroxytryptamine 2 receptor activity and hypercapnia. Am J Physiol 293: R854‐R860, 2007.
 119. Kanbar R, Stornetta RL, Cash DR, Lewis SJ, Guyenet PG. Photostimulation of Phox2b medullary neurons activates cardiorespiratory function in conscious rats. Am J Respir Crit Care Med 182: 1184‐1194, 2010.
 120. Kang BJ, Chang DA, Mackay DD, West GH, Moreira TS, Takakura AC, Gwilt JM, Guyenet PG, Stornetta RL. Central nervous system distribution of the transcription factor Phox2b in the adult rat. J Comp Neurol 503: 627‐641, 2007.
 121. Kawai A, Ballantyne D, Muckenhoff K, Scheid P. Chemosensitive medullary neurones in the brainstem–spinal cord preparation of the neonatal rat. J Physiol 492 (Pt 1): 277‐292, 1996.
 122. Kayaba Y, Nakamura A, Kasuya Y, Ohuchi T, Yanagisawa M, Komuro I, Fukuda Y, Kuwaki T. Attenuated defense response and low basal blood pressure in orexin knockout mice. Am J Physiol Regul Integr Comp Physiol 285: R581‐R593, 2003.
 123. Kiyashchenko LI, Mileykovskiy BY, Maidment N, Lam HA, Wu MF, John J, Peever J, Siegel JM. Release of hypocretin (orexin) during waking and sleep states. J Neurosci 22: 5282‐5286, 2002.
 124. Kogo N, Arita H. In vivo study on medullary H(+)‐sensitive neurons. J Appl Physiol 69: 1408‐1412, 1990.
 125. Kolobow T, Gattinoni L, Tomlinson T, Pierce JE. An alternative to breathing. J Thor Cardiovasc Sur 75: 261‐266, 1978.
 126. Krause KL, Forster HV, Davis SE, Kiner T, Bonis JM, Pan LG, Qian B. Focal acidosis in the pre‐Botzinger complex area of awake goats induces a mild tachypnea. J Appl Physiol 106: 241‐250, 2009.
 127. Kubin L, Tojima H, Davies RO, Pack AI. Serotonergic excitatory drive to hypoglossal motoneurons in the decerebrate cat. Neurosci Lett 139: 243‐248, 1992.
 128. Kuffler SW. Neuroglial cells: Physiological properties and a potassium mediated effect of neuronal activity on the glial membrane potential. Proc R Soc Lond B Biol B 168: 1‐21, 1967.
 129. Kuwaki T. Orexinergic modulation of breathing across vigilance states. Respir Physiol Neurobiol 164: 204‐212, 2008.
 130. Kuwaki T, Li A, Nattie E. State‐dependent central chemoreception: A role of orexin. Respir Physiol Neurobiol 173: 223‐229, 2010.
 131. Kuwaki T, Zhang W, Nakamura A, Deng BS. Emotional and state‐dependent modification of cardiorespiratory function: Role of orexinergic neurons. Auton Neurosci 142: 11‐16, 2008.
 132. Lai YL, Tsuya Y, Hildebrandt J. Ventilatory responses to acute CO2 exposure in the rat. J Appl Physiol 45: 611‐618, 1978.
 133. Lamb TW. Ventilatory responses to intravenous and inspired carbon dioxide in anesthetized cats. Respir Physiol 2: 99‐104, 1966.
 134. Lazarenko RM, Milner TA, Depuy SD, Stornetta RL, West GH, Kievits JA, Bayliss DA, Guyenet PG. Acid sensitivity and ultrastructure of the retrotrapezoid nucleus in Phox2b‐EGFP transgenic mice. J Comp Neurol 517: 69‐86, 2009.
 135. Leusen I. Chemosensitivity of the respiratory center: Influence of CO2 in the cerebral ventricles on respiration. Am J Physiol 176: 390‐444, 1954.
 136. Lewis SM. Awake baboon's ventilatory response to venous and inhaled CO2 loading. J Appl Physiol 39: 417‐422, 1975.
 137. Li A, Emond L, Nattie E. Brainstem catecholaminergic neurons modulate both respiratory and cardiovascular function. Adv Exp Med Biol 605: 371‐376, 2008.
 138. Li A, Nattie E. CO2 dialysis in one chemoreceptor site, the RTN: Stimulus intensity and sensitivity in the awake rat. Respir Physiol Neurobiol 133: 11‐22, 2002.
 139. Li A, Nattie E. Catecholamine neurones in rats modulate sleep, breathing, central chemoreception and breathing variability. J Physiol 570: 385‐396, 2006.
 140. Li A, Nattie E. Serotonin transporter knockout mice have a reduced ventilatory response to hypercapnia (predominantly in males) but not to hypoxia. J Physiol 586: 2321‐2329, 2008.
 141. 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 588: 2935‐2944, 2010.
 142. Li A, Randall M, Nattie EE. CO(2) microdialysis in retrotrapezoid nucleus of the rat increases breathing in wakefulness but not in sleep. J Appl Physiol 87: 910‐919, 1999.
 143. Li A, Zhou S, Nattie E. Simultaneous inhibition of caudal medullary raphe and retrotrapezoid nucleus decreases breathing and the CO2 response in conscious rats. J Physiol 577: 307‐318, 2006.
 144. Linton RA, Miller R, Cameron IR. Ventilatory response to CO2 inhalation and intravenous infusion of hypercapnic blood. Respir Physiol 26: 383‐394, 1976.
 145. Loeschcke HH. Central chemosensitivity and the reaction theory. J Physiol 332: 1‐24, 1982.
 146. Loeschcke HH, Mitchell RA, Katsaros B, Perkins JF, Konig A. Interaction of intracranial chemosensitivity with peripheral afferents to the respiratory centers. Ann N Y Acad Sci 109: 651‐660, 1963.
 147. Lovering AT, Fraigne JJ, Dunin‐Barkowski WL, Vidruk EH, Orem JM. Hypocapnia decreases the amount of rapid eye movement sleep in cats. Sleep 26: 961‐967, 2003.
 148. Lu DC, Erlichman JS, Leiter JC. Diethyl pyrocarbonate (DEPC) inhibits CO2 chemosensitivity in Helix aspersa. Respir Physiol 111: 65‐78, 1998.
 149. Malan A, Wilson TL, Reeves RB. Intracellular pH in cold‐blooded vertebrates as a function of body temperature. Respir Physiol 28: 29‐47, 1976.
 150. Marcus JN, Aschkenasi CJ, Lee CE, Chemelli RM, Saper CB, Yanagisawa M, Elmquist JK. Differential expression of orexin receptors 1 and 2 in the rat brain. J Comp Neurol 435: 6‐25, 2001.
 151. Marina N, Abdala AP, Trapp S, Li A, Nattie EE, Hewinson J, Smith JC, Paton JF, Gourine AV. Essential role of Phox2b‐expressing ventrolateral brainstem neurons in the chemosensory control of inspiration and expiration. J Neurosci 30: 12466‐12473, 2010.
 152. Marjanovic M, Elliott AC, Dawson MJ. The temperature dependence of intracellular pH in isolated frog skeletal muscle: Lessons concerning the Na(+)‐H+ exchanger. J Membr Biol 161: 215‐225, 1998.
 153. Martino PF, Davis S, Opansky C, Krause K, Bonis JM, Czerniak SG, Pan LG, Qian B, Forster HV. Lesions in the cerebellar fastigial nucleus have a small effect on the hyperpnea needed to meet the gas exchange requirements of submaximal exercise. J Appl Physiol 101: 1199‐1206, 2006.
 154. Martino PF, Davis S, Opansky C, Krause K, Bonis JM, Pan LG, Qian B, Forster HV. The cerebellar fastigial nucleus contributes to CO2‐H+ ventilatory sensitivity in awake goats. Respir Physiol Neurobiol 157: 242‐251, 2007.
 155. Martino PF, Hodges MR, Davis S, Opansky C, Pan LG, Krause K, Qian B, Forster HV. CO2/H+ chemoreceptors in the cerebellar fastigial nucleus do not uniformly affect breathing of awake goats. J Appl Physiol 101: 241‐248, 2006.
 156. Martins PJ, D'Almeida V, Pedrazzoli M, Lin L, Mignot E, Tufik S. Increased hypocretin‐1 (orexin‐a) levels in cerebrospinal fluid of rats after short‐term forced activity. Regul Pept 117: 155‐158, 2004.
 157. Meadows GE, Dunroy HM, Morrell MJ, Corfield DR. Hypercapnic cerebral vascular reactivity is decreased, in humans, during sleep compared with wakefulness. J Appl Physiol 94: 2197‐2202, 2003.
 158. Messier ML, Li A, Nattie EE. Inhibition of medullary raphe serotonergic neurons has age‐dependent effects on the CO2 response in newborn piglets. J Appl Physiol 96: 1909‐1919, 2004.
 159. Meza S, Giannouli E, Younes M. Control of breathing during sleep assessed by proportional assist ventilation. J Appl Physiol 84: 3‐12, 1998.
 160. Mitchell RA, Loeschcke HH, Massion WH, Severinghaus JW. Respiratory responses mediated through superficial chemosensitive areas on the medulla. J Appl Physiol 96: 523‐533, 1963.
 161. Mulkey DK, Mistry AM, Guyenet PG, Bayliss DA. Purinergic P2 receptors modulate excitability but do not mediate pH sensitivity of RTN respiratory chemoreceptors. J Neurosci 26: 7230‐7233, 2006.
 162. Mulkey DK, Stornetta RL, Weston MC, Simmons JR, Parker A, Bayliss DA, Guyenet PG. Respiratory control by ventral surface chemoreceptor neurons in rats. Nat Neurosci 7: 1360‐1369, 2004.
 163. Mulkey DK, Wenker IC, Kreneisz O. Current ideas on central chemoreception by neurons and glial cells in the retrotrapezoid nucleus. J Appl Physiol 108: 1433‐1439, 2010.
 164. Nakamura A, Zhang W, Yanagisawa M, Fukuda Y, Kuwaki T. Vigilance state‐dependent attenuation of hypercapnic chemoreflex and exaggerated sleep apnea in orexin knockout mice. J Appl Physiol 102: 241‐248, 2007.
 165. Nambu T, Sakurai T, Mizukami K, Hosoya Y, Yanagisawa M, Goto K. Distribution of orexin neurons in the adult rat brain. Brain Res 827: 243‐260, 1999.
 166. Nattie E. Control and disturbances of cerebrospinal fluid pH. In: Kaila K, Silver RB, editor. pH and Brain Function. New York: Wiley‐Liss, 1998a, p. 629‐650.
 167. Nattie E. Central chemoreceptors, pH and respiratory control. In: Kaila K Silver RB, editor. pH and Brain Function. New York: Wiley‐Liss, 1998b, p. 535‐560.
 168. Nattie E. CO2, brainstem chemoreceptors and breathing. Prog Neurobiol 59: 299‐331, 1999.
 169. Nattie E. Chemoreceptors, pH, and respiratory control. In: Giebisch G SD, editor. The Kidney: Physiology and Pathophysiology. Philadelphia: Lippincott‐Raven, 2001, p. 1983‐1993.
 170. Nattie E, Li A. Muscimol dialysis in the retrotrapezoid nucleus region inhibits breathing in the awake rat. J Appl Physiol 89: 153‐162, 2000.
 171. Nattie E, Li A. Central chemoreception 2005: A brief review. Auton Neurosci 126‐127: 332‐338, 2006a.
 172. Nattie E, Li A. Neurokinin‐1 receptor‐expressing neurons in the ventral medulla are essential for normal central and peripheral chemoreception in the conscious rat. J Appl Physiol 101: 1596‐1606, 2006b.
 173. Nattie E, Li A. Central chemoreception is a complex system function that involves multiple brain stem sites. J Appl Physiol 106: 1464‐1466, 2009.
 174. Nattie E, Li A. Central chemoreception in wakefulness and sleep: Evidence for a distributed network and a role for orexin. J Appl Physiol 108: 1417‐1424, 2010.
 175. Nattie E, Shi J, Li A. Bicuculline dialysis in the retrotrapezoid nucleus (RTN) region stimulates breathing in the awake rat. Respiration physiology 124: 179‐193, 2001.
 176. Nattie EE. Diethyl pyrocarbonate (an imidazole binding substance) inhibits rostral VLM CO2 sensitivity. J Appl Physiol 61: 843‐850, 1986.
 177. Nattie EE. The alphastat hypothesis in respiratory control and acid‐base balance. J Appl Physiol 69: 1201‐1207, 1990.
 178. Nattie EE, Blanchford C, Li A. Retrofacial lesions: Effects on CO2‐sensitive phrenic and sympathetic nerve activity. J Appl Physiol 73: 1317‐1325, 1992.
 179. Nattie EE, Comroe JH Jr. Distinguished lecture of the American Physiological Society Respiration Section: Experimental biology 2010. J Appl Physiol 110: 1‐8, 2011.
 180. Nattie EE, Fung ML, Li A, St John WM. Responses of respiratory modulated and tonic units in the retrotrapezoid nucleus to CO2. Respir Physiol 94: 35‐50, 1993.
 181. Nattie EE, Li A. Retrotrapezoid nucleus lesions decrease phrenic activity and CO2 sensitivity in rats. Respir Physiol 97: 63‐77, 1994.
 182. Nattie EE, Li A. Central chemoreception in the region of the ventral respiratory group in the rat. J Appl Physiol 81: 1987‐1995, 1996.
 183. Nattie EE, Li A. CO2 dialysis in the medullary raphe of the rat increases ventilation in sleep. J Appl Physiol 90: 1247‐1257, 2001.
 184. Nattie EE, Li A. CO2 dialysis in nucleus tractus solitarius region of rat increases ventilation in sleep and wakefulness. J Appl Physiol 92: 2119‐2130, 2002a.
 185. Nattie EE, Li A. Substance P‐saporin lesion of neurons with NK1 receptors in one chemoreceptor site in rats decreases ventilation and chemosensitivity. J Physiol 544: 603‐616, 2002b.
 186. Nattie EE, Li A, Richerson G, Lappi DA. Medullary serotonergic neurones and adjacent neurones that express neurokinin‐1 receptors are both involved in chemoreception in vivo. J Physiol 556: 235‐253, 2004.
 187. Nattie EE, Li AH. Fluorescence location of RVLM kainate microinjections that alter the control of breathing. J Appl Physiol 68: 1157‐1166, 1990.
 188. Nattie EE, Li AH, St John WM. Lesions in retrotrapezoid nucleus decrease ventilatory output in anesthetized or decerebrate cats. J Appl Physiol 71: 1364‐1375, 1991.
 189. Nattie EE, HV Forster. Special issue: Central chemoreception. Respir Physiol & Neurobiol 173: 193‐336, 2010.
 190. Nattie G, Li A. Multiple central chemoreceptor sites: Cell types and function in vivo. Adv Exp Med Biol 605: 343‐347, 2008.
 191. Niblock MM, Gao H, Li A, Jeffress EC, Murphy M, Nattie EE. Fos‐Tau‐LacZ mice reveal sex differences in brainstem c‐fos activation in response to mild carbon dioxide exposure. Brain Res 1311: 51‐63.
 192. Nuding SC, Segers LS, Shannon R, O'Connor R, Morris KF, Lindsey BG. Central and peripheral chemoreceptors evoke distinct responses in simultaneously recorded neurons of the raphe‐pontomedullary respiratory network. Philos Trans R Soc Lond Ser B 364: 2501‐2516, 2009.
 193. Ohtake PJ, Forster HV, Pan LG, Lowry TF, Korducki MJ, Whaley AA. Effects of cooling the ventrolateral medulla on diaphragm activity during NREM sleep. Respir Physiol 104: 127‐135, 1996.
 194. Okada Y, Chen Z, Jiang W, Kuwana S, Eldridge FL. Anatomical arrangement of hypercapnia‐activated cells in the superficial ventral medulla of rats. J Appl Physiol 93: 427‐439, 2002.
 195. Oliven A, Odeh M, Gavriely N. Effect of hypercapnia on upper airway resistance and collapsibility in anesthetized dogs. Respir Physiol 75: 29‐38, 1989.
 196. Olsen ML, Sontheimer H. Functional implications for Kir4.1 channels in glial biology: From K+ buffering to cell differentiation. J Neurochem 107: 589‐601, 2008.
 197. Onimaru H, Ikeda K, Kawakami K. Phox2b, RTN/pFRG neurons and respiratory rhythmogenesis. Respir Physiol Neurobiol 168: 13‐18, 2009.
 198. Orem J. The nature of the wakefulness stimulus for breathing. Prog Clin Biol Res 345: 23‐30; discussion 31, 1990.
 199. Oyamada Y, Ballantyne D, Muckenhoff K, Scheid P. Respiration‐modulated membrane potential and chemosensitivity of locus coeruleus neurones in the in vitro brainstem‐spinal cord of the neonatal rat. J Physiol 513 (Pt 2): 381‐398, 1998.
 200. Pappenheimer JR, Fencl V, Heisey SR, Held D. Role of cerebral fluids in control of respiration as studied in unanesthetized goats. Am J Physiol 208: 436‐450, 1965.
 201. Parisi RA, Neubauer JA, Frank MM, Edelman NH, Santiago TV. Correlation between genioglossal and diaphragmatic responses to hypercapnia during sleep. Am Rev Respir Dis 135: 378‐382, 1987.
 202. Pattyn A, Morin X, Cremer H, Goridis C, Brunet JF. The homeobox gene Phox2b is essential for the development of autonomic neural crest derivatives. Nature 399: 366‐370, 1999.
 203. Penatti EM, Berniker AV, Kereshi B, Cafaro C, Kelly ML, Niblock MM, Gao HG, Kinney HC, Li A, Nattie EE. Ventilatory response to hypercapnia and hypoxia after extensive lesion of medullary serotonergic neurons in newborn conscious piglets. J Appl Physiol 101: 1177‐1188, 2006.
 204. Peyron C, Tighe DK, van den Pol AN, de Lecea L, Heller HC, Sutcliffe JG, Kilduff TS. Neurons containing hypocretin (orexin) project to multiple neuronal systems. J Neurosci 18: 9996‐10015, 1998.
 205. Phillipson EA, Duffin J, Cooper JD. Critical dependence of respiratory rhythmicity on metabolic CO2 load. J Appl Physiol 50: 45‐54, 1981.
 206. Praud JP, Diaz V, Kianicka I, Chevalier JY, Canet E, Thisdale Y. Abolition of breathing rhythmicity in lambs by CO2 unloading in the first hours of life. Respir Physiol 110: 1‐8, 1997.
 207. Putnam RW. CO2 chemoreception in cardiorespiratory control. J Appl Physiol 108: 1796‐1802, 2010.
 208. Ray RS, Corcoran AE, Brust RD, Kim JC, Richerson GB, Nattie E, Dymecki SM. Impaired respiratory and body temperature control upon acute serotonergic neuron inhibition. Science 333: 637‐642, 2011.
 209. Rahn H. Evolution of the gas transport system in vertebrates. Proc R Soc Med 59: 493‐494, 1966.
 210. Rahn H. Why are pH of 7.4 and PCO2 of 40 normal values for man? Bull Eur Physiopathol Respir 12: 5‐13, 1976.
 211. Rahn H, Reeves RB, Howell BJ. Hydrogen ion regulation, temperature, and evolution. Am Rev Respir Dis 112: 165‐172, 1975.
 212. Ransom BR, Sontheimer H. The neurophysiology of glial cells. J Clin Neurophysiol 9: 224‐251, 1992.
 213. Reeves RB. The interaction of body temperature and acid‐base balance in ectothermic vertebrates. Annu Rev Physiol 39: 559‐586, 1977.
 214. Ribas‐Salgueiro JL, Gaytan SP, Crego R, Pasaro R, Ribas J. Highly H+‐sensitive neurons in the caudal ventrolateral medulla of the rat. J Physiol 549: 181‐194, 2003.
 215. Ribas‐Salgueiro JL, Gaytan SP, Ribas J, Pasaro R. Characterization of efferent projections of chemosensitive neurons in the caudal parapyramidal area of the rat brain. Brain Res Bull 66: 235‐248, 2005.
 216. Richerson GB. Response to CO2 of neurons in the rostral ventral medulla in vitro. J Neurophysiol 73: 933‐944, 1995.
 217. Richerson GB. Serotonergic neurons as carbon dioxide sensors that maintain pH homeostasis. Nat Rev Neurosci 5: 449‐461, 2004.
 218. Richerson GB, Wang W, Hodges MR, Dohle CI, Diez‐Sampedro A. Homing in on the specific phenotype(s) of central respiratory chemoreceptors. Exp Physiol 90: 259‐266; discussion 266‐259, 2005.
 219. Richerson GB, Wang W, Tiwari J, Bradley SR. Chemosensitivity of serotonergic neurons in the rostral ventral medulla. Respir Physiol 129: 175‐189, 2001.
 220. Rosin DL, Chang DA, Guyenet PG. Afferent and efferent connections of the rat retrotrapezoid nucleus. J Comp Neurol 499: 64‐89, 2006.
 221. Rosin DL, Weston MC, Sevigny CP, Stornetta RL, Guyenet PG. Hypothalamic orexin (hypocretin) neurons express vesicular glutamate transporters VGLUT1 or VGLUT2. J Comp Neurol 465: 593‐603, 2003.
 222. Rybak IA, Abdala AP, Markin SN, Paton JF, Smith JC. Spatial organization and state‐dependent mechanisms for respiratory rhythm and pattern generation. Prog Brain Res 165: 201‐220, 2007.
 223. Sakurai T. Roles of orexins and orexin receptors in central regulation of feeding behavior and energy homeostasis. CNS Neurol Disord Drug Targets 5: 313‐325, 2006.
 224. Sakurai T. The neural circuit of orexin (hypocretin): Maintaining sleep and wakefulness. Nat Rev Neurosci 8: 171‐181, 2007.
 225. Saper CB, Cano G, Scammell TE. Homeostatic, circadian, and emotional regulation of sleep. J Comp Neurol 493: 92‐98, 2005.
 226. Schlaefke ME, Kille JF, Loeschcke HH. Elimination of central chemosensitivity by coagulation of a bilateral area on the ventral medullary surface in awake cats. Pflugers Arch 378: 231‐241, 1979.
 227. Sears TA, Berger AJ, Phillipson EA. Reciprocal tonic activation of inspiratory and expiratory motoneurones by chemical drives. Nature 299: 728‐730, 1982.
 228. Severinghaus JW. Hans Loeschcke, Robert Mitchell and the medullary CO2 chemoreceptors: A brief historical review. Respir Physiol 114: 17‐24, 1998.
 229. Skatrud JB, Dempsey JA. Interaction of sleep state and chemical stimuli in sustaining rhythmic ventilation. J Appl Physiol 55: 813‐822, 1983.
 230. Smatresk NJ. Chemoreceptor modulation of endogenous respiratory rhythms in vertebrates. Am J Physiol 259: R887‐R897, 1990.
 231. Smith CA, Chenuel BJ, Henderson KS, Dempsey JA. The apneic threshold during non‐REM sleep in dogs: Sensitivity of carotid body vs. central chemoreceptors. J Appl Physiol 103: 578‐586, 2007.
 232. Smith CA, Chenuel BJ, Nakayama H, Dempsey JA. Ventilatory responsiveness to CO2 above & below eupnea: Relative importance of peripheral chemoreception. Adv Exp Med Biol 551: 65‐70, 2004.
 233. Smith CA, Forster HV, Blain GM, Dempsey JA. An interdependent model of central/peripheral chemoreception: Evidence and implications for ventilatory control. Respir Physiol Neurobiol 173: 288‐297, 2010.
 234. Smith CA, Henderson KS, Dempsey JA. Interactive ventilatory effects of carotid body hypoxia and hypocapnia in the unanesthetized dog. Adv Exp Med Biol 393: 313‐316, 1995.
 235. Smith CA, Nakayama H, Dempsey JA. The essential role of carotid body chemoreceptors in sleep apnea. Can J Physiol Pharmacol 81: 774‐779, 2003.
 236. Smith CA, Rodman JR, Chenuel BJ, Henderson KS, Dempsey JA. Response time and sensitivity of the ventilatory response to CO2 in unanesthetized intact dogs: Central vs. peripheral chemoreceptors. J Appl Physiol 100: 13‐19, 2006.
 237. Smith JC, Abdala AP, Koizumi H, Rybak IA, Paton JF. Spatial and functional architecture of the mammalian brain stem respiratory network: A hierarchy of three oscillatory mechanisms. J Neurophysiol 98: 3370‐3387, 2007.
 238. Smith JC, Morrison DE, Ellenberger HH, Otto MR, Feldman JL. Brainstem projections to the major respiratory neuron populations in the medulla of the cat. J Comp Neurol 281: 69‐96, 1989.
 239. Solomon IC. Focal CO2/H +alters phrenic motor output response to chemical stimulation of cat pre‐Botzinger complex in vivo. J Appl Physiol 94: 2151‐2157, 2003.
 240. Somero GN. Proteins and temperature. Annu Rev Physiol 57: 43‐68, 1995.
 241. Spyer KM, Gourine AV. Chemosensory pathways in the brainstem controlling cardiorespiratory activity. Philos Trans R Soc Lond Ser B 364: 2603‐2610, 2009.
 242. St Croix CM, Satoh M, Morgan BJ, Skatrud JB, Dempsey JA. Role of respiratory motor output in within‐breath modulation of muscle sympathetic nerve activity in humans. Circ Res 85: 457‐469, 1999.
 243. St John WM, Glasser RL, King RA. Apneustic breathing after vagotomy in cats with chronic pneumotaxic center lesions. Respir Physiol 12: 239‐250, 1971.
 244. Stornetta RL, Moreira TS, Takakura AC, Kang BJ, Chang DA, West GH, Brunet JF, Mulkey DK, Bayliss DA, Guyenet PG. Expression of Phox2b by brainstem neurons involved in chemosensory integration in the adult rat. J Neurosci 26: 10305‐10314, 2006.
 245. Stornetta RL, Spirovski D, Moreira TS, Takakura AC, West GH, Gwilt JM, Pilowsky PM, Guyenet PG. Galanin is a selective marker of the retrotrapezoid nucleus in rats. J Comp Neurol 512: 373‐383, 2009.
 246. Stunden CE, Filosa JA, Garcia AJ, Dean JB, Putnam RW. Development of in vivo ventilatory and single chemosensitive neuron responses to hypercapnia in rats. Respir Physiol 127: 135‐155, 2001.
 247. Sullivan CE, Murphy E, Kozar LF, Phillipson EA. Waking and ventilatory responses to laryngeal stimulation in sleeping dogs. J Appl Physiol 45: 681‐689, 1978.
 248. Sunanaga J, Deng BS, Zhang W, Kanmura Y, Kuwaki T. CO2 activates orexin‐containing neurons in mice. Respir Physiol Neurobiol 166: 184‐186, 2009.
 249. Sunderram J, Parisi RA, Strobel RJ. Serotonergic stimulation of the genioglossus and the response to nasal continuous positive airway pressure. Am J Respir Crit Care Med 162: 925‐929, 2000.
 250. Szymusiak R, McGinty D. Hypothalamic regulation of sleep and arousal. Ann N Y Acad Sci 1129: 275‐286, 2008.
 251. Takahashi K, Lin JS, Sakai K. Neuronal activity of orexin and non‐orexin waking‐active neurons during wake‐sleep states in the mouse. Neuroscience 153: 860‐870, 2008.
 252. Takakura AC, Moreira TS, Colombari E, West GH, Stornetta RL, Guyenet PG. Peripheral chemoreceptor inputs to retrotrapezoid nucleus (RTN) CO2‐sensitive neurons in rats. J Physiol 572: 503‐523, 2006.
 253. Takakura AC, Moreira TS, Stornetta RL, West GH, Gwilt JM, Guyenet PG. Selective lesion of retrotrapezoid Phox2b‐expressing neurons raises the apnoeic threshold in rats. J Physiol 586: 2975‐2991, 2008.
 254. Taylor NC, Li A, Nattie EE. Medullary serotonergic neurones modulate the ventilatory response to hypercapnia, but not hypoxia in conscious rats. J Physiol 566: 543‐557, 2005.
 255. Taylor NC, Li A, Nattie EE. Ventilatory effects of muscimol microdialysis into the rostral medullary raphe region of conscious rats. Respir Physiol Neurobiol 153: 203‐216, 2006.
 256. Teppema LJ, Veening JG, Kranenburg A, Dahan A, Berkenbosch A, Olievier C. Expression of c‐fos in the rat brainstem after exposure to hypoxia and to normoxic and hyperoxic hypercapnia. J Comp Neurol 388: 169‐190, 1997.
 257. Thannickal TC, Moore RY, Nienhuis R, Ramanathan L, Gulyani S, Aldrich M, Cornford M, Siegel JM. Reduced number of hypocretin neurons in human narcolepsy. Neuron 27: 469‐474, 2000.
 258. Thomas RJ, Daly RW, Weiss JW. Low‐concentration carbon dioxide is an effective adjunct to positive airway pressure in the treatment of refractory mixed central and obstructive sleep‐disordered breathing. Sleep 28: 69‐77, 2005.
 259. Trouth CO, Loeschcke HH, Berndt J. Topography of the respiratory responses to electrical stimulation in the medulla oblongata. Pflugers Arch 339: 153‐170, 1973.
 260. Trzebski A, Kubin L. Is the central inspiratory activity responsible for pCO2‐dependent drive of the sympathetic discharge? J Auton Nerv Syst 3: 401‐420, 1981.
 261. Verin E, Tardif C, Marie JP, Buffet X, Lacoume Y, Delapille P, Pasquis P. Upper airway resistance during progressive hypercapnia and progressive hypoxia in normal awake subjects. Respir Physiol 124: 35‐42, 2001.
 262. Wasserman K, Whipp BJ, Casaburi R, Huntsman DJ, Castagna J, Lugliani R. Regulation of arterial PCO2 during intravenous CO2 loading. J Appl Physiol 38: 651‐656, 1975.
 263. Watanabe S, Kuwaki T, Yanagisawa M, Fukuda Y, Shimoyama M. Persistent pain and stress activate pain‐inhibitory orexin pathways. Neuroreport 16: 5‐8, 2005.
 264. Weese‐Mayer DE, Berry‐Kravis EM, Zhou L, Maher BS, Silvestri JM, Curran ME, Marazita ML. Idiopathic congenital central hypoventilation syndrome: Analysis of genes pertinent to early autonomic nervous system embryologic development and identification of mutations in PHOX2b. Am J Med Genet A 123: 267‐278, 2003.
 265. Weil JV, Byrne‐Quinn E, Sodal IE, Friesen WO, Underhill B, Filley GF, Grover RF. Hypoxic ventilatory drive in normal man. J Clin Invest 49: 1061‐1072, 1970.
 266. Wellman A, Jordan AS, Malhotra A, Fogel RB, Katz ES, Schory K, Edwards JK, White DP. Ventilatory control and airway anatomy in obstructive sleep apnea. Am J Respir Crit Care Med 170: 1225‐1232, 2004.
 267. Wiley RG, Lappi DA. Targeting neurokinin‐1 receptor‐expressing neurons with [Sar9,Met(O2)11 substance P‐saporin. Neurosci Lett 277: 1‐4, 1999.
 268. Williams RH, Jensen LT, Verkhratsky A, Fugger L, Burdakov D. Control of hypothalamic orexin neurons by acid and CO2. Proc Natl Acad Sci U S A 104: 10685‐10690, 2007.
 269. Xie A, Skatrud JB, Barczi SR, Reichmuth K, Morgan BJ, Mont S, Dempsey JA. Influence of cerebral blood flow on breathing stability. J Appl Physiol 106: 850‐856, 2009.
 270. Xie A, Skatrud JB, Morgan B, Chenuel B, Khayat R, Reichmuth K, Lin J, Dempsey JA. Influence of cerebrovascular function on the hypercapnic ventilatory response in healthy humans. J Physiol 577: 319‐329, 2006.
 271. Xu F, Zhang Z, Frazier DT. Microinjection of acetazolamide into the fastigial nucleus augments respiratory output in the rat. J Appl Physiol 91: 2342‐2350, 2001.
 272. Yamamoto WS, Edwards MW Jr. Homeostasis of carbon dioxide during intravenous infusion of carbon dioxide. J Appl Physiol 15: 807‐818, 1960.
 273. Yoshida Y, Fujiki N, Nakajima T, Ripley B, Matsumura H, Yoneda H, Mignot E, Nishino S. 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: 1075‐1081, 2001.
 274. Young JK, Wu M, Manaye KF, Kc P, Allard JS, Mack SO, Haxhiu MA. Orexin stimulates breathing via medullary and spinal pathways. J Appl Physiol 98: 1387‐1395, 2005.
 275. Zhang W, Shimoyama M, Fukuda Y, Kuwaki T. Multiple components of the defense response depend on orexin: Evidence from orexin knockout mice and orexin neuron‐ablated mice. Auton Neurosci 126‐127: 139‐145, 2006.

Related Articles:

Acid‐Base Balance in Cerebral Fluids
Brain Blood Flow and Control of Breathing
Control of Breathing During Sleep
Gas Exchange in Acid‐Base Disturbances
Peripheral Chemoreceptors and Their Sensory Neurons in Chronic States of Hypo‐ and Hyperoxygenation
Peripheral Chemoreceptors: Function and Plasticity of the Carotid Body
Neural Control of the Upper Airway: Integrative Physiological Mechanisms and Relevance for Sleep Disordered Breathing
Control of Breathing in the Fetus and the Newborn
Genetic Diseases: Congenital Central Hypoventilation, Rett, and Prader‐Willi Syndromes
Central Sleep Apnea

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

Eugene Nattie, Aihua Li. Central Chemoreceptors: Locations and Functions. Compr Physiol 2012, 2: 221-254. doi: 10.1002/cphy.c100083