Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2008 Jul 1.
Published in final edited form as: Respir Physiol Neurobiol. 2007 Jan 20;157(1):154–161. doi: 10.1016/j.resp.2007.01.009

The influence of chronic hypoxia upon chemoreception

Frank L Powell 1
PMCID: PMC1964780  NIHMSID: NIHMS23102  PMID: 17291837

Abstract

Carotid body chemoreceptors are essential for time-dependent changes in ventilatory control during chronic hypoxia. Early theories of ventilatory acclimatization to hypoxia focused on time-dependent changes in known ventilatory stimuli, such as small changes in arterial pH that may play a significant role in some species. However, plasticity in the cellular and molecular mechanisms of carotid body chemoreception play a major role in ventilatory acclimatization to hypoxia in all species studied. Chronic hypoxia causes changes in (a) ion channels (potassium, sodium, calcium) to increase glomus cell excitability, and (b) neurotransmitters (dopamine, acetylcholine, ATP) and neuromodulators (endothelin-1) to increase carotid body afferent activity for a given PO2 and optimize O2-sensitivity. O2-sensing heme-containing molecules in the carotid body have not been studied in chronic hypoxia. Plasticity in medullary respiratory centers processing carotid body afferent input also contributes to ventilatory acclimatization to hypoxia. It is not known if the same mechanisms occur in patients with chronic hypoxemia from lung disease or high altitude natives.

Keywords: carotid body, hypoxic ventilatory response, neural plasticity, ventilatory acclimatization

1. Introduction

Exposure to environmental hypoxia elicits chemoreflexes that increase ventilation and tend to reduce the effects of decreased inspired PO2 on arterial PO2. With chronic hypoxia (hours to months), the increase in ventilation is greater than that found during acute exposure (minutes) to the same degree of hypoxia and this further minimizes the impact of the environmental challenge by raising arterial PO2 (Weil, 1986). This time-dependent increase in ventilation with chronic hypoxia is known as ventilatory acclimatization to hypoxia (VAH) and involves plasticity in the chemoreflexes that control ventilation (Powell et al., 1998).

Most of what we know about the neurobiological mechanisms of the acute ventilatory response to hypoxia is based on studies of the carotid body and the same is true for plasticity in ventilatory chemoreflexes with chronic hypoxia. Using a variety of experimental preparations, from conscious large animals to in vitro co-cultures of chemoreceptor cells and afferent neurons, several laboratories have demonstrated physiologically significant increases in O2-sensitivity of carotid bodies with chronic hypoxia (described below). Chronic hypoxia also induces plasticity in central respiratory centers that process afferent input from carotid bodies (Dwinell and Powell, 1999) and alters other aspects of the neural control of breathing, such as control of arterial PCO2 (Weil, 1986; Smith et al., 2001). However, this paper focuses on plasticity in the carotid bodies with chronic sustained hypoxia and their effects on breathing. It also considers how important carotid body plasticity is relative to other mechanisms of VAH. The effects of chronic intermittent hypoxia on carotid bodies (Prabhakar, 2001) are not covered here.

2. Carotid body chemoreceptors

Several excellent reviews are available on carotid body structure and function (Fidone and Gonzalez, 1986; Gonzalez et al., 1994; Gonzalez et al., 1995). The carotid body is a complex sensory organ that is richly perfused by arterial blood and innervated by a branch of the glossopharyngeal nerve (i.e. carotid sinus nerve), which makes it ideally suited for monitoring arterial blood gases and pH. It also receives sympathetic innervation which may control blood flow and other functions. Carotid bodies are composed of two main cell types called glomus cells and sustentacular cells. Glomus cells, also known as Type I cells, are of neural crest origin and are generally believed to be the primary chemoreceptors. Sustentacular cells, or Type II cells, are similar to glial cells but their function in the carotid body is not known. In response to changes in arterial PO2, PCO2 and pH, glomus cells presumably release an excitatory neurotransmitter to depolarize carotid sinus nerve afferent terminals. This sends action potentials to respiratory centers in the CNS and stimulates ventilation. Hence, the first question to answer for understanding plasticity of the carotid body ventilatory reflex during chronic hypoxia is how the physiological stimuli for arterial chemoreceptors change under such conditions.

2.1 Chemoreceptor stimuli in chronic hypoxia

Early theories and experiments for ventilatory acclimatization to hypoxia focused on potential changes in the known stimuli for chemoreceptors and assumed sensory transduction was unchanged. For example, one of the original theories for ventilatory acclimatization to hypoxia focused on changes in the pH of cerebrospinal fluid (CSF pH) as a stimulus to central CO2- sensitive chemoreceptors. While the change in CO2 regulation in chronic hypoxia remains unexplained, it is now generally accepted that time-dependent changes in CSF pH do not explain ventilatory acclimatization (Forster and Dempsey, 1981). Failure of the CSF pH theories to explain ventilatory acclimatization to hypoxia lead to the idea of neural plasticity, i.e. central nervous system (CNS) processing of afferent information was enhanced by chronic hypoxia (Forster et al., 1974). Today we know that plasticity occurs in both the arterial chemoreceptors and the CNS, as detailed later in this paper.

In general, carotid body stimulation is predicted to decrease during chronic hypoxia. The time dependent increase in ventilation during acclimatization acts to decrease PCO2 further and raise PO2 back towards control levels (Rahn and Otis, 1949). On average, arterial pH does not change with chronic hypoxia relative to the value measured during acute hypoxia (Dempsey and Forster, 1982). In other words, metabolic compensation is complete for the time-dependent decrease in PaCO2 with increased ventilation during acclimatization to hypoxia. Therefore in general, none of the arterial stimuli for carotid bodies are increasing so they cannot explain time-dependent increases in ventilation. However metabolic compensation is more effective in some experimental animals, for example decreasing pHa by 0.05 units relative to the acute hypoxic value in chronically hypoxic rats (Olson, Jr. and Dempsey, 1978) and cats (Barnard et al., 1987). The consequences of this are considered at the end of the next section.

2.2. Afferent activity from carotid bodies during chronic hypoxia

Chronic hypoxia has been shown to increase O2-sensitivity in the carotid body and increase the frequency of action potentials in the carotid sinus nerve for a given PO2 in several different experimental preparations. Continuous recording of the carotid sinus nerve in goats for 5 hrs of hypoxia show steady increases in the frequency of action potentials (Nielsen et al., 1988) while this does not occur with 4 hrs exposure to hypercapnia. Single unit recordings from cats exposed to 28 days of hypoxia (Barnard et al., 1987), as well as whole carotid sinus nerve recordings from cats exposed to 48 hours of hypoxia (Vizek et al., 1987), also show increased O2-sensitivity. In vitro carotid body preparations from rats acclimatized to hypoxia from 3 to 16 days show increased O2-sensitivity too (Chen et al., 2002b). All of these studies find about a 100% increase in the frequency of action potentials in acute hypoxia following chronic hypoxia.

Chronic hypoxia does not affect carotid body sensitivity to pH (Barnard et al., 1987) but in some animals there may be a small increase in pH stimulation of carotid bodies (see above). Data from normoxic cats can be used to predict that a 0.05 drop in pH (~5 nmol/L) would increase action potential frequency in the carotid sinus nerve by only 10% (Lahiri and DeLaney, 1975). However, the same study shows that such a change in pHa and carotid body activity could almost double ventilation, which is as great as the effect of decreasing PaO2 from 95 to 40 Torr (Fig. 9 in (Lahiri and DeLaney, 1975). Hence the role of a decrease in arterial pH could be considerable in ventilatory acclimatization to hypoxia but experiments to quantify this effect have not been done. In any case, the predicted effects of such small changes in pHa on carotid body afferent activity is small compared to the effect of chronic hypoxia described above.

3. Cellular mechanisms of chemoreception

Functional plasticity involves the molecular mechanisms of chemoreception, ion channels and neurotransmitters in the carotid body (recently reviewed by (Bisgard, 2000) and (Dinger et al., 2003)). There are important histological changes in carotid bodies during chronic hypoxia, such as glomus cell hypertrophy and decreased covering of glomus cells by sustentacular cells (Kusakabe et al., 1993). The latter change increases the potential area available for gap junction connections between glomus cells, which have been shown to enhance glomus cell sensitivity (Eyzaguirre and Abudara, 1999). However, these morphological changes generally take much longer than the increased O2-sensitivity of the carotid body described above (days versus hours) so they do not appear necessary for functional plasticity in chemoreception.

3.1 Molecular chemoreceptors

Studies on isolated carotid bodies in vitro and isolated glomus cells in culture indicate that the glomus cells contain the primary O2-sensing mechanism in carotid bodies (Fidone and Gonzalez, 1986; Gonzalez et al., 1994). The metabolic hypothesis of chemoreception proposes specialized forms of heme-containing molecules, which are generally involved in oxygen metabolism, are the primary sensors of hypoxia (Prabhakar, 2006). An early version of the metabolic hypotheses was based on a low-O2 affinity cytochrome a3 found in carotid bodies and presumably in the mitochondria (Mills and Jöbsis, 1972). More recent experiments show mitochondria in carotid bodies depolarize at O2 tensions much higher than those causing changes in other cell types, although this may represent modulation by factors such as nitric oxide (NO) instead of unique heme molecules (Donnelly and Carroll, 2005). Experimental evidence also exists for NADPH oxidase and heme oxygenase 2 (HO-2) being involved in O2-sensing (Prabhakar and Jacono, 2005).

Several links between mitochondrial function and glomus cell depolarization and excitatory neurotransmitter release have been proposed. However, ATP levels, as indicators of cellular energy state do not correlate well with glomus cell activation (Donnelly and Carroll, 2005). Release of mitochondrial calcium stores in carotid bodies was hypothesized to be a key step in O2-transduction (Duchen and Biscoe, 1992), although this may not be simply increasing intracellular calcium in glomus cells to promote excitatory neurotransmitter release. Donnelly and Carroll (Donnelly and Carroll, 2005) reviewed the literature and conclude that “the calcium buffering function of mitochondria through the mitochondrial potential and the ability of the mitochondrial redox state to modulate other cell processes through generation of reactive oxygen intermediates” may be important for O2 sensing.

There is relatively little information on the effects of chronic hypoxia on these putative molecular mechanisms of O2-sensing in the carotid body during chronic hypoxia. However, recent work on the pulmonary circulation has demonstrated a role for NADPH oxidase in the pathogenesis of pulmonary hypertension during chronic hypoxia (Liu et al., 2005). Mice without a subunit protein of the superoxide producing NADPH oxidase (gp91phox) have significantly reduced levels of superoxide production and reduced vasoconstrictor responses to endothelin-1 in chronic hypoxia compared to wild type mice. Under normal conditions, gp91phox is not necessary for O2-sensing by carotid bodies (Roy et al., 2000) but NADPH oxidase may play an important role in plasticity during chronic hypoxia when it produces more superoxide that could enhance the effects of other neuromodulators. As discussed later, endothelin-1 appears to be involved in increased carotid body O2-sensitivity with chronic hypoxia.

3.2 Ion channels

Glomus cells have ion channels that are both sensitive and not sensitive to hypoxia. However, these ion channels interact, for example through voltage or calcium dependency, to determine glomus cell excitability and the response to acute hypoxia. Chronic hypoxia has been shown to affect both kinds of ion channels with the net result being increased excitability.

O2-sensitive potassium (K+) channels in glomus cells include (a) TASK-like background channels, which also are sensitive to pH and determine resting potentials, and (b) voltage gated K+ channels (Kv) that decrease conductance at very modest levels of hypoxia (PO2 = 80 Torr), which is consistent with chemoreceptor function (Patel and Honore, 2001). Chronic hypoxia is reported to decrease K+ channel density in glomus cells from chronically hypoxic neonatal rats and this would increase excitability (Hempleman, 1995). Chronic hypoxia is reported to increase expression and current density of Ca2+ activated K+ channels (maxi-K channels) in a HEK293 cell culture system (Hartness et al., 2003). However, it is difficult to extrapolate between organs or model cell systems because regulatory proteins required for O2-sensitivity and Kv subunits are differentially expressed in different cell types. (Patel and Honore, 2001).

Although sodium (Na+) channels are expressed at relatively low levels in glomus cells (e.g. 10% of the level found in the squid giant axon), they are important for glomus cell membrane and action potentials (Gonzalez et al., 1994). Na+ channel density increases in glomus cells cultured in chronic hypoxia, which could contribute to increased excitability of glomus cells after stimulation by hypoxia (Stea et al., 1992). (Hempleman, 1995) found increased Na+ currents in glomus cells from chronically hypoxic neonatal rat pups, which could increase excitability given the decrease he found in K+ channel density in the same preparation (see above).

(Hempleman, 1996) also found increased calcium channel density in glomus cells from chronically hypoxic neonatal rats. This was primarily the L-type Ca2+ channel, which is enhanced by hypoxia more than other types of Ca2+ channels found in glomus cells, i.e. P/Q, N and R or resistant types (Prabhakar and Jacono, 2005).

There is also some evidence that the afferent nerve endings of the carotid sinus nerve have O2-sensitive ion channels. Donnelly recently showed that a persistent Na+ current (INaP), which is sensitive to riluzole and tetrodotoxin and present only on afferent nerves, plays a critical role in carotid body O2-sensitivity (Faustino and Donnelly, 2006). Chemosensitivity in the afferent nerve endings is an attractive hypothesis that might explain experimental results such as the recovery of chemosensitivity in the carotid sinus nerve after removal of the carotid body (Mitchell et al., 1972), and the large variation in effects of a given neurotransmitter or neuromodulators on chemoreception between species (see below). This hypothesis deserves further study because it has been difficult to experimentally test given the challenges in making neural recordings from the fine nerve endings in the carotid body (Hayashida et al., 1980). However, at this time the mechanism is only hypothetical and any role for afferent nerve chemosensitivity in plasticity of carotid body function during chronic hypoxia remains hypothetical.

3.3. Neurochemicals

Dopamine (DA) is one of the most abundant neurotransmitters in the carotid body and it has been studied extensively, yet its function still is not completely understood. DA is synthesized by tyrosine hydroxylase (TH) in glomus cells and it is released from glomus cells by physiological chemoreceptor stimuli (Gonzalez et al., 1994). Despite such biochemical evidence for DA being an excitatory neurotransmitter in carotid bodies, pharmacological and physiological studies indicate that DA is primarily inhibitory (Fidone et al., 1990). D2 dopamine receptors (D2-R) are present on both afferent nerve terminals and glomus cells and these are generally inhibitory. The physiological role of DA in the carotid body appears to be primarily an auto-regulatory mechanism that limits glomus cell depolarization during chemoreceptor stimulation.

Given the inhibitory effects of DA in the carotid body, it was hypothesized that decreased DA effects may explain increased O2-sensitivity in chronically hypoxic carotid bodies (Bisgard, 2000). However, such dis-inhibition would have to overcome effects of increased DA levels in the carotid body with chronic hypoxia, which results from effects of hypoxic inducible transcription factor (HIF-1α) on TH (see below). Experiments on cats exposed to 2 days of chronic hypoxia showed that D2-R antagonists completely reversed the time-dependent increase in ventilatory O2-sensitivity that occurred during VAH (Tatsumi et al., 1995). This supported the idea that decreased inhibition by DA at D2-R explains VAH. However, subsequent studies in rats showed that while decreased inhibition by DA in the carotid body occurred after 2 days of chronic hypoxia, the inhibitory effects of DA were restored after 7 days of chronic hypoxia (Huey et al., 2000). Hence, it appears that the primary function of DA in the carotid, and the physiological significance of changing DA and D2-R during chronic hypoxia, is to maintain normal O2-sensitivity in the face of long-term changes of stimulation and base-line activity levels. Re-establishing control levels of sensitivity after exposure to chronic perturbations appears to be a general principle of sensory systems (Turrigiano, 1999). Also, it is consistent with the proposal that a “push-pull” mechanism between excitatory and inhibitory neurotransmitters in carotid bodies may be more important in chronic hypoxia compared to “normal” conditions (Prabhakar, 2006).

Acetylcholine (ACh) has been identified as a classical excitatory neurotransmitter in the carotid body and may play a role in increased O2-sensitivity with chronic hypoxia (reviewed by (Bisgard, 2000; Kim et al., 2004; Prabhakar and Jacono, 2005). Nicotinic cholinergic receptors on carotid body afferent nerves are increased by chronic hypoxia and the excitatory effect of exogenous ACh on the carotid body is enhanced by chronic hypoxia. However, nicotinic receptor antagonists do not reverse the enhanced response to acute hypoxia after acclimatization suggesting another transmitter that is co-released with ACh may be important.

ATP, which is co-released with ACh from glomus cells, has been shown recently to make an important contribution to increased O2-sensitivity of carotid bodies after chronic hypoxia (He et al., 2006). Blocking the P2X2 purinoreceptors decreased the neural response to acute hypoxia in carotid bodies from both normoxic control and chronically hypoxic rats. Also, the absolute decrease in neural activity was greater in chronically hypoxic preparations consistent with ATP contributing to enhanced O2-sensitivity. However, P2X2 blockaded did not completely eliminate the response to hypoxia in either condition indicating that ATP acts in concert with other excitatory mechanisms in the carotid body. This could be an excitatory neurotransmitter, such as ACh discussed above, or perhaps enhanced electrical coupling between glomus cells and afferent nerve terminals considering that chronic hypoxia up-regulated gap junction proteins (connexin-43) in carotid bodies (Chen et al., 2002a).

Endothelin 1 (ET-1) has the strongest evidence for explaining increased O2-sensitivity in carotid bodies with chronic hypoxia. ET-1 is a vasoactive peptide that increases with chronic hypoxia in O2-sensitive tissues such as the pulmonary artery (Li et al., 1994) and carotid bodies (Chen et al., 2002b). ET-1 alone does not excite carotid bodies but it does enhance the response to hypoxia and this effect is mediated by ETA receptors on glomus cells (Chen et al., 2000). Chronic hypoxia increases the expression of ET-1 and ETA receptors in glomus cells and blocking ETA receptors reverses the increased hypoxic sensitivity observed in chronic hypoxia, returning it to control levels (Chen et al., 2002b). The effect of ET-1 to enhance hypoxic sensitivity of glomus cells involves increased calcium currents that could result from ion channel modulation by signals from ETA receptors (Chen et al., 2002c).

Other neurotransmitters present in the carotid body and postulated to be important in chemoreception, namely norepinephrine and serotonin, have not been shown to play an important role in acclimatization of the carotid body to date (reviewed by (Bisgard, 2000). Changes in NO in the carotid body during chronic hypoxia are predicted to inhibit O2-sensitivity (reviewed by (Bisgard, 2000). However, NO might be a modulator that optimizes the ability of the carotid body to respond to changes in O2 level at different chronic levels of O2, similar to DA discussed above.

3.4 Hypoxic sensitive gene expression

Many of the cellular changes in carotid body chemoreception during chronic hypoxia result from changes in gene expression. A myriad of transcription factors are sensitive to oxygen (Cummins and Taylor, 2005) but only a few have been studied in the carotid body (e.g. CREB, AP-1, p53). Hypoxic inducible factor 1a (HIF-1α) has been studied the most in the carotid body (Kline et al., 2002; Wilson et al., 2005). HIF-1α is increased by chronic hypoxia in brainstem nuclei important for the ventilatory control (Soulage et al., 2004; Soliz et al., 2005) suggesting it may play a role in CNS, as well as carotid body, plasticity in chronic hypoxia. It is important to note that plasticity in carotid bodies or the CNS need not necessarily result from hypoxic sensitive gene expression because the tonic changes in activity levels with chronic hypoxic stimulation could also induce gene expression.

4. Plasticity in the carotid body ventilatory reflex

Hypoxic stimulation of carotid body chemoreceptors leads to a powerful reflex increase in ventilation known as the hypoxic ventilatory response (HVR). Given the enhanced O2-sensitivity in carotid chemoreceptors during chronic hypoxia described above, it is not surprising that hours to months cause changes in the control of breathing (Weil, 1986; Smith et al., 2001).

4.1 Chronic hypoxia increases the HVR

The increased HVR observed in hours to months of chronic hypoxia is opposite the decreased, or blunted, HVR seen in high altitude natives (see below). There has been some disagreement about the effects of chronic hypoxia on the HVR in both human and animal studies but it appears this can be explained by changes in CO2 levels during the acute hypoxic challenges used to measure the HVR. If the HVR is tested in isocapnia, so PaCO2 not allowed to fall with hyperventilation during acute hypoxia, then the HVR is increased by chronic hypoxia (Powell et al., 2000a). The quantitative roles of different mechanisms in explaining the increased HVR in chronic hypoxia remain to be established, however. There may be some effects from small changes in arterial pH (see above) but the major determinants appear to be plasticity in carotid bodies and the CNS.

4.2 Carotid body denervation experiments

Carotid body denervation experiments have established a definitive role for arterial chemoreceptors in ventilatory acclimatization to hypoxia. Bilateral section of the carotid sinus nerve or excision of the carotid bodies causes hypoventilation and increased PaCO2 normoxia, and eliminates or significantly reduces the acute ventilatory response to hypoxia or carotid body stimuli, e.g. sodium cyanide (Forster, 2003). Carotid body denervation in cats, dogs, goats, ponies, rabbits, rats and sheep eliminates the time-dependent increases in ventilation normally observed in chronic hypoxia (Bouverot et al., 1973; Vizek et al., 1987; Olson, Jr. et al., 1988; Smith et al., 2001).

It is also relevant that ventilatory acclimatization to hypoxia does not occur in animals that exhibit some recovery of their acute HVR long after carotid body denervation (Forster, 2003). Plasticity in the HVR following carotid body denervation has been observed in several species but it is not due to reinnervation. Hypothesized mechanisms include upregulation of alternate sensory pathways (e.g. serotonin-mediated aortic body chemoreception) or upregulation of the efferent limb of a reflex (e.g. serotonin-mediated facilitation of phrenic motor neurons). However, even after the acute HVR and normoxic PaCO2 levels have returned towards normal in dogs, goats, ponies, and rabbits with denervated carotid bodies, acclimatization to chronic hypoxia remains attenuated or even eliminated. This is not surprising considering the plasticity in O2-chemoreception in carotid bodies described above. However, it cannot be used to quantify the role of peripheral (carotid body) versus central (CNS) plasticity in normal acclimatization to hypoxia (Dempsey and Forster, 1982). For example, CNS plasticity may not occur in carotid body denervated animals because the tonic increase in afferent input in chronic hypoxia would not occur.

4.3 Plasticity in carotid body

Increased O2-sensitivity of the carotid body with chronic hypoxia is a major determinant of the increased HVR with chronic hypoxia. This was demonstrated by Bisgard and his colleagues using goats with the carotid bodies isolated from the systemic circulation. This preparation allows PO2 or PCO2 in blood perfusing the carotid body to be changed independently from systemic arterial PO2or PCO2 for up to eight hours. Six hours of isolated carotid body hypoxia caused a progressive increase in ventilation above the acute hypoxic ventilatory response (Busch et al., 1985). Such time-dependent changes in ventilation did not occur with (a) 6 hours of carotid body hypercapnia (Bisgard et al., 1986a), (b) 4 hr of systemic hypocapnia (Bisgard et al., 1986b) or, (c) systemic arterial hypoxia (Weizhen et al., 1992). These results show the response is specific to carotid body hypoxia and not the result of increased carotid body stimulation or activity, or changes in PO2 or pH at central chemoreceptors and respiratory centers in the brain. Plasticity induced by decreased PO2 in the carotid body during the first hours of chronic hypoxia appears sufficient to explain ventilatory acclimatization for up to 6 hrs of hypoxia in goats. However, these results do not preclude other mechanisms of plasticity (e.g. in the CNS) occurring during longer hypoxic exposures, as discussed below.

4.4 Plasticity in the Central Nervous System (CNS)

The idea of chronic hypoxia inducing plasticity in the CNS to explain changes in the arterial chemoreflex actually precedes the idea of plasticity in the carotid body. Experiments in the 1970s were not finding support for the CSF pH theory of ventilatory acclimatization to hypoxia and altered responsiveness of medullary ventilatory control centers was postulated as an alternative explanation (Dempsey and Forster, 1982). Experimental evidence for this alternative included an increased ventilatory response to doxapram, a chemoreceptor stimulant, in subjects acclimatized to high altitude for 2 to 3 weeks (Forster et al., 1974). Assuming the effect of doxapram on carotid bodies is unchanged by chronic hypoxia, an increased ventilatory response to a similar dose of doxapram implies a greater ventilatory motor output for a comparable chemoreceptor afferent input to the CNS. Hence, such enhanced responsiveness of CNS respiratory centers can be described as an increase in the CNS gain of the HVR (Powell et al., 2000b).

More definitive evidence for an increased CNS gain of the HVR with chronic hypoxia was obtained from experiments measuring phrenic nerve activity in response to electrical stimulation of the carotid sinus nerve in anesthetized rats (Dwinell and Powell, 1999). Seven days of chronic hypoxia significantly increased the CNS gain of the HVR and there was a non-significant tendency for it to increase after only two days of hypoxia. In contrast, the physiological indicators of ventilatory acclimatization to hypoxia, i.e. metabolic rate, ventilation and arterial PCO2, do not change between 2 and 7 days of hypoxia in rats (Powell et al., 2000b). Therefore, experiments demonstrating a specific mechanism of plasticity acting at a specific time point in chronic hypoxia, may not apply with shorter or longer exposures. Hence, the demonstration of increased carotid body O2-sensitivity with no change in CNS gain of the HVR in cats after 48 hours of hypoxia (Vizek et al., 1987), and similar ventilation and blood gases after 48 hours or longer exposures (Tatsumi et al., 1991), does not rule out increased CNS gain of the HVR at later times. Similar logic applies to the experiments showing ventilatory acclimatization in goats after 4 to 8 hours of hypoxia can be explained by increased carotid body O2-sensitivity (Busch et al., 1985); experiments have not been performed that would disprove central plasticity with longer hypoxic exposures in other animals.

There are also chemosensitive areas in the brain that respond directly to O2 (Neubauer and Sunderram, 2004) and CO2 (Nattie, 2006). As mentioned above, central CO2-chemoreceptors especially have been of great interest for changes in ventilatory control with chronic hypoxia. However, studies to date focused on changes in CSF pH as a stimulus and recent studies have not supported a mechanism involving central chemoreceptors for ventilatory acclimatization (Smith et al., 2001). Plasticity in either central O2 or CO2 chemoreceptors could play a role in control of breathing changes during chronic hypoxia but this has not been studied.

It is also possible that chronic hypoxia could affect interactions between central and arterial chemoreflexes. There is recent evidence for functional and anatomical links between the carotid body ventilatory chemoreflex and CO2-sensitive areas in the retrotrapezoid nucleus (Takakura et al., 2006). Also, caudal regions of the nucleus of the solitary tract is the site of both the primary synapse from carotid sinus nerve afferents (Gonzalez et al., 1994) and CO2 chemosensitivity that stimulates ventilation (Dean et al., 1990; Coates et al., 1993). However, most evidence indicates the interaction between central and peripheral chemoreception is additive (Cunningham et al., 1986). Hence, changes in central chemoreceptive drive would not be predicted to increase the slope of the HVR, which is observed in chronic hypoxia. There is some evidence for a multiplicative interaction between central and peripheral ventilatory chemoreflexes in humans but it is not large (Robbins, 1988). However, increasing central chemoreceptive drive actually decreased the slope of the isocapnic HVR in the normoxic control rats while chronic hypoxia increased the slope (Powell et al., 2000a).

5. Physiological significance of carotid body plasticity in hypoxia

The increased HVR observed during acclimatization to hypoxia is physiologically significant by increasing ventilation, which reduces the impact of environmental hypoxia on oxygen levels in the body. However, longer exposures to hypoxia at high altitude for years to generations “blunts” the HVR (Weil et al., 1971). The blunted HVR is a very general response observed in birds and mammals that are native to high altitudes (Black and Tenney, 1980; Weil, 1986), although not in all human populations (Beall et al., 1997; Leon-Velarde and Richalet, 2006). This suggests a blunted HVR may result from natural selection and be an advantage in hypoxia at altitude. It can be speculated that a blunted HVR is advantageous by reducing the oxygen cost of breathing only if there are other adaptations. For example, Andean high altitude natives have more efficient metabolic processes in hypoxia compared to lowlanders (Rupert and Hochachka, 2001). Further research is necessary to determine if a blunted HVR is advantageous during very long-term hypoxia.

The significance of carotid body plasticity for chronic hypoxemia with disease is not clear either. Although much of the basic research done on this topic uses the rationale that it will help us understand heart and lung disease and devise better treatments, we still do not know if plasticity occurs in the carotid bodies and chemoreflexes of chronically hypoxemic patients. This is because the problem is extremely difficult to study in patients with chronic lung disease for several reasons. First, the ventilatory chemoreflexes are extremely variable between individuals but the genetic basis for this is unknown. Therefore, it is impossible to estimate the healthy normal ventilatory response in an individual before he or she had the disease. Also, there are no longitudinal studies of the problem in populations susceptible to chronic lung disease to overcome this limitation to date. Second, individual patients with lung disease may have different degrees of abnormality in gas exchange and lung mechanics. Such changes in the efferent arm of ventilatory chemoreflexes can have very powerful effects on the physiological control of arterial blood gases and this is difficult to distinguish from problems with the sensory or integrative components of the reflex. For example, Younes (1995) pointed out that ventilation can increase enough to overcome a six-fold increase in respiratory resistance if ventilation is stimulated by an increase in PaCO2 that is too small to reliably measure. Finally, it is difficult to determine if specific mechanisms described for experimental animals (e.g. plasticity in the afferent vs. CNS integrating components of the HVR) occur in patients because of limitations in interpreting the results from reflex studies in humans (Powell et al., 2000a). Hence, the significance of plasticity in the HVR for lung disease remains to be determined also.

6. Conclusions

Chronic hypoxia causes plasticity in O2-sensitive chemoreceptors and the hypoxic ventilatory response reflex. Ventilatory acclimatization to hours to months of hypoxia is a robust and universal response that reduces the impact of chronic hypoxia on oxygen levels in the body. Yet acclimatization results from a wide variety of species-specific molecular and neurochemical mechanisms that operate over very different time domains. Some of these mechanisms may be redundant. Hence, carotid body plasticity with chronic hypoxia is consistent with the “chemosome” hypothesis that proposes multiple O2-sensors to explain the wide range of O2 sensitivity in the carotid body (Prabhakar, 2006). Similarly, multiple levels of plasticity in the arterial chemoreflex (i.e. both sensory and central integrating components) appear to have evolved in concert so ventilatory sensitivity to hypoxia is optimized for the prevailing level of environmental oxygen.

Acknowledgments

Supported by NIH HL 81823.

Footnotes

Submitted for a Special Issue of Respiratory Physiology & Neurobiology ‘Sensing hypoxia: carotid body mechanisms and reflexes in health and disease’, edited by P. Kumar and N.R Prabhakar

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  1. Barnard P, Andronikou S, Pokorski M, Smatresk N, Mokashi A, Lahiri S. Time-dependent effect of hypoxia on carotid body chemosensory function. J Appl Physiol. 1987;63:685–691. doi: 10.1152/jappl.1987.63.2.685. [DOI] [PubMed] [Google Scholar]
  2. Bisgard GE. Carotid body mechanisms in acclimatization to hypoxia. Respir Physiol. 2000;121:237–246. doi: 10.1016/s0034-5687(00)00131-6. [DOI] [PubMed] [Google Scholar]
  3. Bisgard GE, Busch MA, Daristotle L, Berssenbrugge AD, Forster HV. Carotid body hypercapnia does not elicit ventilatory acclimatization in goats. Respir Physiol. 1986a;65:113–125. doi: 10.1016/0034-5687(86)90010-1. [DOI] [PubMed] [Google Scholar]
  4. Bisgard GE, Busch MA, Forster HV. Ventilatory acclimatization to hypoxia is not dependent on cerebral hypocapnic alkalosis. J Appl Physiol. 1986b;60:1011–1015. doi: 10.1152/jappl.1986.60.3.1011. [DOI] [PubMed] [Google Scholar]
  5. Black CP, Tenney SM. Oxygen transport during progressive hypoxia in high-altitude and sea-level waterfowl. Respir Physiol. 1980;39:217–239. doi: 10.1016/0034-5687(80)90046-8. [DOI] [PubMed] [Google Scholar]
  6. Bouverot P, Candas V, Libert JP. Role of the arterial chemoreceptors in ventilatory adaptation to hypoxia of awake dogs and rabbits. Respir Physiol. 1973;17:209–219. doi: 10.1016/0034-5687(73)90062-5. [DOI] [PubMed] [Google Scholar]
  7. Busch MA, Bisgard GE, Forster HV. Ventilatory acclimatization to hypoxia is not dependent on arterial hypoxemia. J Appl Physiol. 1985;58(6):1874–1880. doi: 10.1152/jappl.1985.58.6.1874. [DOI] [PubMed] [Google Scholar]
  8. Chen J, He L, Dinger B, Fidone S. Cellular mechanisms involved in rabbit carotid body excitation elicited by endothelin peptides. Respir Physiol. 2000;121:13–23. doi: 10.1016/s0034-5687(00)00113-4. [DOI] [PubMed] [Google Scholar]
  9. Chen J, He L, Dinger B, Stensaas L, Fidone S. Chronic hypoxia upregulates connexin43 expression in rat carotid body and petrosal ganglion. J Appl Physiol. 2002a;92:1480–1486. doi: 10.1152/japplphysiol.00077.2001. [DOI] [PubMed] [Google Scholar]
  10. Chen J, He L, Dinger B, Stensaas L, Fidone S. Role of endothelin and endothelin Atype receptor in adaptation of the carotid body to chronic hypoxia. Am J Physiol Lung Cell Mol Physiol. 2002b;282:L1314–L1323. doi: 10.1152/ajplung.00454.2001. [DOI] [PubMed] [Google Scholar]
  11. Chen Y, Tipoe GL, Liong E, Leung S, Lam SY, Iwase R, Tjong YW, Fung ML. Chronic hypoxia enhances endothelin-1-induced intracellular calcium elevation in rat carotid body chemoreceptors and up-regulates ETA receptor expression. Pflugers Arch. 2002c;443:565–573. doi: 10.1007/s00424-001-0728-2. [DOI] [PubMed] [Google Scholar]
  12. Coates EL, Li A, Nattie EE. Widespread sites of brain stem ventilatory chemoreceptors. J Appl Physiol. 1993;75:5–14. doi: 10.1152/jappl.1993.75.1.5. [DOI] [PubMed] [Google Scholar]
  13. Cummins EP, Taylor CT. Hypoxia-responsive transcription factors. Pflugers Arch. 2005;450:363–371. doi: 10.1007/s00424-005-1413-7. [DOI] [PubMed] [Google Scholar]
  14. Cunningham DJC, Robbins PA, Wolff CB. Integration of respiratory responses to changes in alveolar partial pressure of CO2 and O2 and in arterial pH. In: Cherniack NS, Widdicombe JG, editors. Handbook of Physiology: The Respiratory System - Control of Breathing. Waverly Press, Inc; Baltimore, MD: 1986. pp. 475–528. [Google Scholar]
  15. Dean JB, Bayliss DA, Erickson JT, Lawing WL, Millhorn DE. Depolarization and stimulation of neurons in nucleus tractus solitarii by carbon dioxide does not require chemical synaptic input. Neuroscience. 1990;36:207–216. doi: 10.1016/0306-4522(90)90363-9. [DOI] [PubMed] [Google Scholar]
  16. Dempsey JA, Forster HV. Mediation of ventilatory adaptations. Physiol Rev. 1982;62(1):262–346. doi: 10.1152/physrev.1982.62.1.262. [DOI] [PubMed] [Google Scholar]
  17. Dinger B, He L, Chen J, Stensaas L, Fidone S. Mechanisms of morphological and functional plasticity in the chronically hypoxic carotid body. In: Lahiri S, Semenza G, Prabhakar NR, editors. Oxygen Sensing: Responses and Adaptation to Hypoxia. Marcel Dekker; New York: 2003. pp. 439–465. [Google Scholar]
  18. Donnelly DF, Carroll JL. Mitochondrial function and carotid body transduction. High Alt Med Biol. 2005;6:121–132. doi: 10.1089/ham.2005.6.121. [DOI] [PubMed] [Google Scholar]
  19. Duchen MR, Biscoe TJ. Mitochondrial function in type I cells isolated from rabbit arterial chemoreceptors. J Physiol. 1992;450:13–31. doi: 10.1113/jphysiol.1992.sp019114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Dwinell MR, Powell FL. Chronic hypoxia enhances the phrenic nerve response to arterial chemoreceptor stimulation in anesthetized rats. J Appl Physiol. 1999;87:817–823. doi: 10.1152/jappl.1999.87.2.817. [DOI] [PubMed] [Google Scholar]
  21. Eyzaguirre C, Abudara V. Carotid body glomus cells: chemical secretion and transmission (modulation?) across cell-nerve ending junctions. Respir Physiol. 1999;115:135–149. doi: 10.1016/s0034-5687(99)00020-1. [DOI] [PubMed] [Google Scholar]
  22. Faustino EV, Donnelly DF. An important functional role of persistent Na+ current in carotid body hypoxia transduction. J Appl Physiol. 2006;101:1076–1084. doi: 10.1152/japplphysiol.00090.2006. [DOI] [PubMed] [Google Scholar]
  23. Fidone SJ, Gonzalez C. Initiation and control of chemoreceptor activity in the carotid body. In: Cherniack NS, Widdicombe JG, editors. Handbook of Physiology: The Respiratory System - Control of Breathing. Waverly Press, Inc; Baltimore, MD: 1986. pp. 247–312. [Google Scholar]
  24. Fidone SJ, Gonzalez C, Obeso A, Gomez-Nino A, Dinger B. Biogenic amine and neuropeptide transmitters in carotid body chemotransmission: experimental findings and perspectives. In: Sutton JR, Coates G, Remmers JE, editors. Hypoxia: The Adaptations. BC Decker Inc; Toronto: 1990. pp. 116–126. [Google Scholar]
  25. Forster HV. Plasticity in the control of breathing following sensory denervation. J Appl Physiol. 2003;94:784–794. doi: 10.1152/japplphysiol.00602.2002. [DOI] [PubMed] [Google Scholar]
  26. Forster HV, Dempsey JA. Ventilatory Adaptations. In: Hornbein TF, editor. Regulation of Breathing. Marcel Dekker, Inc; New York: 1981. pp. 845–904. [Google Scholar]
  27. Forster HV, Dempsey JA, Vidruk EH, Do Pico G. Evidence of altered regulation of ventilation during exposure to hypoxia. Respir Physiol. 1974;20:379–392. doi: 10.1016/0034-5687(74)90034-6. [DOI] [PubMed] [Google Scholar]
  28. Gonzalez C, Almaraz L, Obeso A, Rigual R. Carotid body chemoreceptors: from natural stimuli to sensory discharges. Physiol Rev. 1994;74:829–898. doi: 10.1152/physrev.1994.74.4.829. [DOI] [PubMed] [Google Scholar]
  29. Gonzalez C, Dinger BG, Fidone SJ. Mechanisms of carotid body chemoreception. In: Dempsey JA, Pack AI, editors. Regulation of Breathing. Marcel Dekker, Inc; New York, Basel, Hong Kong: 1995. pp. 391–471. [Google Scholar]
  30. Hartness ME, Brazier SP, Peers C, Bateson AN, Ashford ML, Kemp PJ. Post-transcriptional control of human maxiK potassium channel activity and acute oxygen sensitivity by chronic hypoxia. J Biol Chem. 2003;278:51422–51432. doi: 10.1074/jbc.M309463200. [DOI] [PubMed] [Google Scholar]
  31. Hayashida Y, Koyano H, Eyzaguirre C. An intracellular study of chemosensory fibers and endings. J Neurophysiol. 1980;44(6):1077–1088. doi: 10.1152/jn.1980.44.6.1077. [DOI] [PubMed] [Google Scholar]
  32. He L, Chen J, Dinger B, Stensaas L, Fidone S. Effect of chronic hypoxia on purinergic synaptic transmission in rat carotid body. J Appl Physiol. 2006;100:157–162. doi: 10.1152/japplphysiol.00859.2005. [DOI] [PubMed] [Google Scholar]
  33. Hempleman SC. Sodium and potassium currents in neonatal rat carotid body cells following in vivo chronic hypoxia. Brain Res. 1995;699:42–50. doi: 10.1016/0006-8993(95)00850-p. [DOI] [PubMed] [Google Scholar]
  34. Hempleman SC. Increased calcium currents in carotid body glomus cells following in vivo acclimatization to chronic hypoxia. J Neurophysiol. 1996;76:1880–1886. doi: 10.1152/jn.1996.76.3.1880. [DOI] [PubMed] [Google Scholar]
  35. Huey KA, Brown IP, Jordan MC, Powell FL. Changes in dopamine D2-receptor modulation of the hypoxic ventilatory response with chronic hypoxia. Respir Physiol. 2000;123:177–187. doi: 10.1016/s0034-5687(00)00175-4. [DOI] [PubMed] [Google Scholar]
  36. Kim DK, Prabhakar NR, Kumar GK. Acetylcholine release from the carotid body by hypoxia: evidence for the involvement of autoinhibitory receptors. J Appl Physiol. 2004;96:376–383. doi: 10.1152/japplphysiol.00726.2003. [DOI] [PubMed] [Google Scholar]
  37. Kline DD, Peng YJ, Manalo DJ, Semenza GL, Prabhakar NR. Defective carotid body function and impaired ventilatory responses to chronic hypoxia in mice partially deficient for hypoxia-inducible factor 1 alpha. Proc Natl Acad Sci U S A. 2002;99:821–826. doi: 10.1073/pnas.022634199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Kusakabe T, Powell FL, Ellisman MH. Ultrastructure of the glomus cells in the carotid body of chronically hypoxic rats: with a special reference to the similarity of the amphibian glomus cells. Anat Rec. 1993;237:220–227. doi: 10.1002/ar.1092370209. [DOI] [PubMed] [Google Scholar]
  39. Lahiri S, DeLaney RG. Relationship between carotid chemoreceptor activity and ventilation in the cat. Respir Physiol. 1975;24:267–286. doi: 10.1016/0034-5687(75)90018-3. [DOI] [PubMed] [Google Scholar]
  40. Li H, Chen SJ, Chen YF, Meng QC, Durand J, Oparil S, Elton TS. Enhanced endothelin-1 and endothelin receptor gene expression in chronic hypoxia. J Appl Physiol. 1994;77:1451–1459. doi: 10.1152/jappl.1994.77.3.1451. [DOI] [PubMed] [Google Scholar]
  41. Liu JQ, Erbynn EM, Folz RJ. Chronic hypoxia-enhanced murine pulmonary vasoconstriction: role of superoxide and gp91phox. Chest. 2005;128:594S–596S. doi: 10.1378/chest.128.6_suppl.594S. [DOI] [PubMed] [Google Scholar]
  42. Mills E, Jöbsis FF. Mitochondrial respiratory chain of carotid body and chemoreceptor response to changes in oxygen tension. J Neurophysiol. 1972;5:405–428. doi: 10.1152/jn.1972.35.4.405. [DOI] [PubMed] [Google Scholar]
  43. Mitchell RA, Sinha AK, McDonald DM. Chemoreceptive properties of regenerated endings of the carotid sinus nerve. Brain Res. 1972;43:681–685. doi: 10.1016/0006-8993(72)90430-1. [DOI] [PubMed] [Google Scholar]
  44. Nattie EE. Central chemoreception 2005: a brief review. Autonomic Neuroscience. 2006;126-127:332–338. doi: 10.1016/j.autneu.2006.02.003. [DOI] [PubMed] [Google Scholar]
  45. Neubauer JA, Sunderram J. Oxygen-sensing neurons in the central nervous system. J Appl Physiol. 2004;96:367–374. doi: 10.1152/japplphysiol.00831.2003. [DOI] [PubMed] [Google Scholar]
  46. Nielsen AM, Bisgard GE, Vidruk EH. Carotid chemoreceptor activity during acute and sustained hypoxia in goats. J Appl Physiol. 1988;65(4):1796–1802. doi: 10.1152/jappl.1988.65.4.1796. [DOI] [PubMed] [Google Scholar]
  47. Olson EB, Jr, Dempsey JA. Rat as a model for human like ventilatory adaptation to chronic hypoxia. J Appl Physiol. 1978;44:763–769. doi: 10.1152/jappl.1978.44.5.763. [DOI] [PubMed] [Google Scholar]
  48. Olson EB, Jr, Vidruk EH, Dempsey JA. Carotid body excision significantly changes ventilatory control in awake rats. J Appl Physiol. 1988;64:666–671. doi: 10.1152/jappl.1988.64.2.666. [DOI] [PubMed] [Google Scholar]
  49. Patel AJ, Honore E. Molecular physiology of oxygen-sensitive potassium channels. Eur Respir J. 2001;18:221–227. doi: 10.1183/09031936.01.00204001. [DOI] [PubMed] [Google Scholar]
  50. Powell FL, Dwinell MR, Aaron EA. Measuring ventilatory acclimatization to hypoxia: comparative aspects. Respir Physiol. 2000a;122:271–284. doi: 10.1016/s0034-5687(00)00165-1. [DOI] [PubMed] [Google Scholar]
  51. Powell FL, Huey KA, Dwinell MR. Central nervous system mechanisms of ventilatory acclimatization to hypoxia. Respir Physiol. 2000b;121:223–236. doi: 10.1016/s0034-5687(00)00130-4. [DOI] [PubMed] [Google Scholar]
  52. Powell FL, Milsom WK, Mitchell GS. Time domains of the hypoxic ventilatory response. Respir Physiol. 1998;112:123–134. doi: 10.1016/s0034-5687(98)00026-7. [DOI] [PubMed] [Google Scholar]
  53. Prabhakar NR. Oxygen sensing during intermittent hypoxia: cellular and molecular mechanisms. J Appl Physiol. 2001;90:1986–1994. doi: 10.1152/jappl.2001.90.5.1986. [DOI] [PubMed] [Google Scholar]
  54. Prabhakar NR. O2 sensing at the mammalian carotid body: why multiple O2 sensors and multiple transmitters? Exp Physiol. 2006;91:17–23. doi: 10.1113/expphysiol.2005.031922. [DOI] [PubMed] [Google Scholar]
  55. Prabhakar NR, Jacono FJ. Cellular and molecular mechanisms associated with carotid body adaptations to chronic hypoxia. High Alt Med Biol. 2005;6:112–120. doi: 10.1089/ham.2005.6.112. [DOI] [PubMed] [Google Scholar]
  56. Rahn H, Otis AB. Man’s respiratory response during and after acclimatization to high altitude. Am J Physiol. 1949;157:445–462. doi: 10.1152/ajplegacy.1949.157.3.445. [DOI] [PubMed] [Google Scholar]
  57. Robbins PA. Evidence for interaction between the contributions to ventilation from the central and peripheral chemoreceptors in man. J Physiol. 1988;401:503–518. doi: 10.1113/jphysiol.1988.sp017175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Roy A, Rozanov C, Mokashi A, Daudu P, Al mehdi AB, Shams H, Lahiri S. Mice lacking in gp91 phox subunit of NAD(P)H oxidase showed glomus cell [Ca2+]i and respiratory responses to hypoxia. Brain Res. 2000;872:188–193. doi: 10.1016/s0006-8993(00)02458-6. [DOI] [PubMed] [Google Scholar]
  59. Rupert JL, Hochachka PW. The evidence for hereditary factors contributing to high altitude adaptation in Andan natives: A review. High Alt Med Biol. 2001;2:235–256. doi: 10.1089/152702901750265332. [DOI] [PubMed] [Google Scholar]
  60. Smith CA, Dempsey JA, Hornbein TF. Control of breathing at high altitude. In: Hornbein TF, Schoene RB, editors. High Altitude. Marcel Dekker, Inc; New York: 2001. pp. 139–173. [Google Scholar]
  61. Soliz J, Joseph V, Soulage C, Becskei C, Vogel J, Pequignot JM, Ogunshola O, Gassmann M. Erythropoietin regulates hypoxic ventilation in mice by interacting with brainstem and carotid bodies. J Physiol. 2005;568:559–571. doi: 10.1113/jphysiol.2005.093328. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Soulage C, Pascual O, Roux JC, Denavit-Saubie M, Pequignot JM. Chemosensory inputs and neural remodeling in carotid body and brainstem catecholaminergic cells. Adv Exp Med Biol. 2004;551:53–58. doi: 10.1007/0-387-27023-x_9. [DOI] [PubMed] [Google Scholar]
  63. Stea A, Jackson A, Nurse CA. Hypoxia and N6, 02′ -dibutyryladenosine 3′,5′ -cyclic monophosphate, but not nerve growth factor, induce Na+ channels and hypertrophy in chromaffin-like arterial chemoreceptors. Proc Natl Acad Sci. 1992;89:9469–9473. doi: 10.1073/pnas.89.20.9469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. 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. 2006;572:503–523. doi: 10.1113/jphysiol.2005.103788. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Tatsumi K, Pickett CK, Weil JV. Attenuated carotid body hypoxic sensitivity after prolonged hypoxic exposure. J Appl Physiol. 1991;70:748–755. doi: 10.1152/jappl.1991.70.2.748. [DOI] [PubMed] [Google Scholar]
  66. Tatsumi K, Pickett CK, Weil JV. Possible role of dopamine in ventilatory acclimitization to high altitude. Resp Physiol. 1995;99:63–73. doi: 10.1016/0034-5687(94)00084-d. [DOI] [PubMed] [Google Scholar]
  67. Turrigiano GG. Homeostatic plasticity in neuronal networks: the more things change, the more they stay the same. Trends Neurosci. 1999;22:221–227. doi: 10.1016/s0166-2236(98)01341-1. [DOI] [PubMed] [Google Scholar]
  68. Vizek M, Pickett CK, Weil JV. Increased carotid body hypoxic sensitivity during acclimatization to hypobaric hypoxia. J Appl Physiol. 1987;63:2403–2410. doi: 10.1152/jappl.1987.63.6.2403. [DOI] [PubMed] [Google Scholar]
  69. Weil JV. Ventilatory control at high altitude. In: Cherniack NS, Widdicombe JG, editors. Handbook of Physiology: The Respiratory System - Control of Breathing. American Physiological Society; Bethesda: 1986. pp. 703–728. [Google Scholar]
  70. Weil JV, Byrne-Quinn E, Sodal IE, Filley GF, Grover RF. Acquired attenuation of chemoreceptor function in chronically hypoxic man at high altitude. J Clin Invest. 1971;50:186–195. doi: 10.1172/JCI106472. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Weizhen N, Engwall MJA, Daristotle L, Pizzaro J, Bisgard GE. Ventilatory effects of prolonged systemic (CNS) hypoxia in awake goats. Respir Physiol. 1992;87:37–48. doi: 10.1016/0034-5687(92)90098-h. [DOI] [PubMed] [Google Scholar]
  72. Wilson DF, Roy A, Lahiri S. Immediate and long-term responses of the carotid body to high altitude. High Alt Med Biol. 2005;6:97–111. doi: 10.1089/ham.2005.6.97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Younes M. Mechanisms of respiratory load compensation. In: Dempsey JA, Pack AI, editors. Regulation of Breathing. Marcel Dekker, Inc; New York, Basel, Hong Kong: 1995. pp. 867–922. [Google Scholar]

RESOURCES