Skip to main content
The Journal of Physiology logoLink to The Journal of Physiology
. 2006 May 11;574(Pt 1):113–123. doi: 10.1113/jphysiol.2006.108381

AMP-activated protein kinase and the regulation of Ca2+ signalling in O2-sensing cells

A Mark Evans 1
PMCID: PMC1817783  PMID: 16709639

Abstract

All cells respond to metabolic stress. However, a variety of specialized cells, commonly referred to as O2-sensing cells, are acutely sensitive to relatively small changes in PO2. Within a variety of organisms such O2-sensing cells have evolved as vital homeostatic mechanisms that monitor O2 supply and alter respiratory and circulatory function, as well as the capacity of the blood to transport O2. Thereby, arterial PO2 may be maintained within physiological limits. In mammals, for example, two key tissues that contribute to this process are the pulmonary arteries and the carotid bodies. Constriction of pulmonary arteries by hypoxia optimizes ventilation–perfusion matching in the lung, whilst carotid body excitation by hypoxia initiates corrective changes in breathing patterns via increased sensory afferent discharge to the brain stem. Despite extensive investigation, the precise mechanism(s) by which hypoxia mediates these responses has remained elusive. It is clear, however, that hypoxia inhibits mitochondrial function in O2-sensing cells over a range of PO2 that has no such effect on other cell types. This raised the possibility that AMP-activated protein kinase might function to couple mitochondrial oxidative phosphorylation to Ca2+ signalling mechanisms in O2-sensing cells and thereby underpin pulmonary artery constriction and carotid body excitation by hypoxia. Our recent investigations have provided significant evidence in support of this view.

Introduction

Although the mechanism of chemotransduction by hypoxia is keenly debated, a general consensus is developing as to the processes that underpin pulmonary artery constriction and carotid body excitation by hypoxia, respectively.

Hypoxic pulmonary vasoconstriction (HPV) is the critical and distinguishing characteristic of the arteries that feed the lung (von Euler & Liljestrand, 1946). In marked contrast, systemic arteries dilate in response to hypoxia to meet the metabolic demands of the tissues they supply (Roy & Sherrington, 1890). Physiologically, HPV contributes to ventilatation–perfusion matching in the lung by diverting blood flow to O2-rich areas (Fishman, 1976; Voelkel, 1986). However, when alveolar hypoxia is global, as in diseases such as emphysema and cystic fibrosis, HPV leads to hypoxic pulmonary hypertension (HPH) and right heart failure (Voelkel, 1986). HPV is driven by the response to hypoxia of two different cell types, namely the pulmonary arterial smooth muscle and endothelial cells.

HPV had been presumed to be mediated by inhibition of voltage-gated K+ channel (Kv) currents, subsequent membrane depolarization and voltage-gated Ca2+ influx (Michelakis et al. 2004). However, it is now clear that constriction by hypoxia of pulmonary artery rings may be induced in the absence of extracellular Ca2+, i.e. under conditions where voltage-gated influx is abolished (Dipp et al. 2001). Consistent with this fact, a growing body of evidence now suggests that acute HPV is primarily initiated and maintained by the release of Ca2+ from smooth muscle sarcoplasmic reticulum (SR) stores via ryanodine receptors (RyRs; Salvaterra & Goldman, 1993; Jabr et al. 1997; Dipp & Evans, 2001; Dipp et al. 2001; Morio & McMurtry, 2002) and by consequent Ca2+ influx via a store-refilling current (Kang et al. 2003; Ng et al. 2005; Wang et al. 2005). This process of SR Ca2+ release by hypoxia is promoted and maintained by cyclic adenosine diphosphate ribose (cADPR; (Dipp & Evans, 2001; Wilson et al. 2001; Aley et al. 2005, 2006), a Ca2+ mobilizing pyridine nucleotide that activates RyRs (Lee et al. 1989; Rusinko & Lee, 1989; Galione et al. 1991; Walseth et al. 1991), and is a prerequisite for the full expression of HPV (Dipp & Evans, 2001). Subsequently, smooth muscle constriction is augmented by myofilament Ca2+ senstization that is likely evoked by an endothelium-derived vasoconstrictor (Demiryurek et al. 1991; Kovitz et al. 1993; Robertson et al. 1995; Gaine et al. 1998; Dipp & Evans, 2001; Dipp et al. 2001; Robertson et al. 2001; Wilson et al. 2001), the release of which appears dependent on transmembrane Ca2+ influx into pulmonary artery endothelial cells (Dipp et al. 2001).

Carotid body excitation by hypoxia serves a discrete physiological function and is initiated by a different mechanism. Here carotid body type I cells, otherwise referred to as glomus cells, monitor systemic arterial PO2 and initiate corrective changes in breathing patterns (Heymans et al. 1930). Upon exposure to hypoxia, voltage-gated Ca2+ influx into type I cells initiates neurosecretion (Fidone et al. 1988; Gonzalez et al. 1994; Chen et al. 1997) and thereby increases sensory afferent discharge to the brain stem. This process is driven by the inhibition of what have been termed O2-sensitive K+ channels (Lopez-Barneo et al. 1988; Peers, 1990; Stea & Nurse, 1991; Lopez-Lopez & Gonzalez, 1992; Wyatt & Peers, 1992, 1995; Buckler, 1997, 1999; Buckler et al. 2000). Thus, membrane depolarization is triggered, and in turn activates voltage-gated Ca2+ channels rather than ER Ca2+ release (Buckler & Vaughan-Jones, 1994). The molecular identity of the O2-sensitive K+ channel has, however, been contested and there are clear differences between species (Peers, 1990; Ganfornina & Lopez-Barneo, 1991, 1992; Buckler et al. 2000; Sanchez et al. 2002; Lopez-Lopez et al. 2003). Despite this fact, it is now generally accepted that hypoxia selectively inhibits TASK-like K+ channels (Buckler et al. 2000) and large conductance Ca2+-activated (BKCa) K+ channels (Wyatt & Peers, 1995) in rat carotid body type I cells and thereby triggers carotid body excitation, whilst inhibition of Kv channels is thought to underpin depolarization in mouse and rabbit carotid body type I cells (Ganfornina & Lopez-Barneo, 1992; Sanchez et al. 2002; Perez-Garcia et al. 2004).

Mitochondrial oxidative phosphorylation and the regulation by hypoxia of O2-sensing cells

The precise mechanism(s) by which hypoxia elicits the aforementioned responses represents a more contentious issue. It has been suggested that O2-sensitive signal transduction pathways independent of mitochondria may play a role (Youngson et al. 1993; Prabhakar, 1998; Prabhakar & Overholt, 2000; Williams et al. 2004). However, the only consistent finding is that relatively mild hypoxia inhibits mitochondrial oxidative phosphorylation and it has been suggested that this may underpin, at least in part, cell activation (Mills & Jobsis, 1972; Rounds & McMurtry, 1981; Duchen & Biscoe, 1992a, b; Wyatt & Buckler, 2004). Thus, depolarization of the mitochondrial membrane potential and/or an increase in β-NAD(P)H levels has been reported in all O2-sensing cells studied to date (Archer et al. 1986; Duchen & Biscoe, 1992a, b; Youngson et al. 1993; Shigemori et al. 1996; Leach et al. 2001) and over a range of PO2 that elicits no such response in other cell types (Duchen & Biscoe, 1992b). Others have argued that the affinity of cytochrome c oxidase for O2 is too high to allow for the inhibition of mitochondrial oxidative phosphorylation by physiological hypoxia, despite the above and the fact that respiratory control by O2 may be determined by cell- and tissue-specific differences in (1) the metabolic environment of mitochondria, (2) the O2 affinity of cytochrome c oxidase, and (3) the dependence of the O2 affinity of cytochrome c oxidase on metabolic state and rate (Gnaiger et al. 1998). Such controversy has been fuelled by the fact that the mechanism by which inhibition of mitochondrial oxidative phosphorylation may couple to discrete and cell-specific Ca2+ signalling mechanisms has remained elusive. Previous investigations have focused on the possible role of the cellular energy status (ATP; Rounds & McMurtry, 1981; Buescher et al. 1991; Leach et al. 1998, 2000), reduced redox couples (Archer et al. 1993; Wilson et al. 2001) and reactive O2 species (Killilea et al. 2000; Leach et al. 2001; Waypa et al. 2001; Prabhakar & Kumar, 2004), respectively, but extensive investigation of these hypotheses has delivered conflicting data and failed to unite the field since its inception in 1930 (Gonzalez et al. 2002; He et al. 2005). However, previous assessments of the role of the cellular energy status may have been limited by knowledge of the identity of the energy variable that might signal the response (Buescher et al. 1991); significantly, the role of the AMP/ATP ratio had not been considered until recently.

Regulation of AMP-activated protein kinase by metabolic stresses

Over the past decade AMP-activated protein kinase (AMPK), a serine threonine kinase comprising a catalytic α subunit and regulatory β and γ subunits, has come to prominence as a metabolic fuel gauge which monitors the cellular AMP/ATP ratio as an index of metabolic stress. AMPK is activated in response to a variety of metabolic stresses that either increase cellular ATP consumption or reduce ATP supply via mitochondrial oxidative phosphorylation, for example heat shock and metabolic poisons in hepatocytes (Corton et al. 1995; Shaw et al. 2005), exercise in skeletal muscle (Winder & Hardie, 1996; Sakamoto et al. 2005, 2006) and ischaemia in the heart (Marsin et al. 2000; Sakamoto et al. 2006). Activation of AMPK is consequent to the action of adenylate kinase, which catalyses the reaction that converts 2 molecules of ADP to ATP + AMP in order to counter, in the short term, any reduction in ATP supply. Consequently, the cellular ADP/ATP ratio is converted into a much larger increase in the AMP/ATP ratio, which varies as the square of the ADP/ATP ratio (Hardie & Hawley, 2001). Subsequent binding of AMP to two Bateman domains on the γ subunit of AMPK (Hawley et al. 1995; Scott et al. 2002; Kemp, 2004) triggers activation of the kinase by: (1) allosteric regulation via the γ subunit, (2) permitting phosphorylation of the α subunit at Thr-172 by an upstream kinase that is a complex between the tumour suppressor kinase LKB1 and two accessory proteins STRAD and MO25 (Hawley et al. 2003; Woods et al. 2003; Shaw et al. 2004, 2005), and (3) inhibiting dephosphorylation of AMPK. In the absence of metabolic stress, each of these processes is antagonized by high concentrations of ATP, for which the Bateman domains on the γ subunit have lower affinity than they do for AMP. Thus, AMPK is regulated by a triple mechanism that is exquisitely sensitive to very small changes in the AMP/ATP ratio (Hardie & Hawley, 2001). It should be noted, however, that other upstream kinases may regulate AMPK. For example, recent studies have shown that calmodulin-dependent kinase kinases, especially CaMKKβ, may also activate AMPK via the phosphorylation of Thr-172. This does not require an increase in the AMP/ATP ratio, but is instead promoted by a rise in intracellular Ca2+ concentration (Anderson et al. 1998; Hawley et al. 2005; Hurley et al. 2005; Woods et al. 2005). It is possible therefore that CaMKKβ may provide positive feedback in response to the rise in intracellular Ca2+ induced by AMPK in O2-sensing cells. In this respect it may be significant that expression of the two CaMKK isoforms (α and β) is mainly restricted to cells of neural origin (Anderson et al. 1998), such as carotid body type I cells.

Upon activation, AMPK promotes catabolic pathways in order to maintain ATP supply, whilst switching off non-essential ATP-consuming (anabolic) pathways (see Hardie et al. 2006). Thus, the primary targets for AMPK had been presumed to be genes and proteins involved in energy metabolism, particularly lipid and carbohydrate metabolism, and little consideration was given to the role of AMPK as a mediator of physiological processes, despite the fact that AMPK is a serine threonine kinase. However, it is now recognized that AMPK can also target non-metabolic processes (Hardie, 2005), and hence the focus of this issue of The Journal of Physiology. In this respect it is important to note that while AMPK is ubiquitously expressed throughout eukaryotic cells, at least 12 different heterotrimers may be formed from multiple isoforms (Cheung et al. 2000) of the catalytic α subunit (α1 and α2) and regulatory β (β1 and β2) and γ (γ1–3) subunits, splice variants of which may add to the diversity. Thus, the selective expression of a particular AMPK isozyme(s) could determine, at least in part, both cell-and system-specific responses to metabolic stresses. Taking the above into account and the fact that inhibition of mitochondrial oxidative phosphorylation by hypoxia will promote a rise in the AMP/ATP ratio, I proposed that AMPK might act as a mediator of chemotransduction by hypoxia in O2-sensing cells (Evans, 2004). This was a teleological argument based on the premise that (1) the evolution of complex physiological systems would utilize successful processes in different ways as long as they were fit for purpose, and (2) the observation that discrete populations of cells have evolved to serve as O2 sensors in a variety of species. For example, in common with its effects on mammalian pulmonary arteries and carotid body type I cells, hypoxia constricts systemic arteries that lead from the gills of cyclostomes such as the Pacific hagfish (Olson et al. 2001), constricts arteries that feed the skin of the amphibia (Malvin & Walker, 2001), and activates neuroepithelial cells in the gill of teleosts (Jonz et al. 2004).

Is a specific AMPK subunit isoform combination expressed in pulmonary arterial smooth muscle?

In 1946, von Euler and Liljestrand demonstrated that hypoxia without hypercapnia induced constriction within the pulmonary circulation and proposed that HPV might aid ventilation–perfusion matching, by diverting blood flow away from poorly ventilated areas of the lung (von Euler & Liljestrand, 1946). This is now recognized as the critical and distinguishing characteristic of pulmonary arteries; systemic arteries dilate in response to tissue hypoxaemia (Roy & Sherrington, 1890). Clearly therefore one would expect a mediator of HPV to offer in some way the pulmonary selectivity required to elicit this response. One determining factor could be the nature of the AMPK isozyme(s) expressed in pulmonary arterial smooth muscle. Thus, we characterized the subunit isoforms present. These investigations showed that the α1 and α2 catalytic subunit isoforms and the β2, γ1, γ2 and γ3 regulatory subunit isoforms are expressed in smooth muscle from the second and third order branches of the pulmonary arterial tree, although the γ3 isoform accounts for such a small fraction of the total AMPK activity that its functional relevance is open to question. Thus, it is most likely that four AMPK subunit isoform combinations contribute to the regulation of pulmonary arterial smooth muscle, namely α1β2γ1, α1β2γ2, α2β2γ1 and α2β2γ2. However, the catalytic activity was mainly accounted for by the α1 subunit (> 80%) whilst 60% was associated with the regulatory γ1 subunit, consistent with the α1β2γ1 heterotrimer being predominant (Evans et al. 2005, 2006a, b). Intriguingly, the level of AMPK activities in the second and third order branches of the pulmonary arterial tree contrasted markedly with the activities in the main pulmonary artery that feeds the lung and systemic (mesenteric) arteries. Briefly, in smooth muscle from the main pulmonary artery the activity associated with the α1 catalytic subunit isoform alone declined by 50%. Thus, AMPK-α1 activity is inversely related to pulmonary artery diameter, as is the magnitude of pulmonary artery constriction by hypoxia (Kato & Staub, 1966) and, significantly, the enzyme activities for the synthesis and metabolism of cADPR, a primary trigger for SR Ca2+ release by hypoxia and HPV (Dipp & Evans, 2001; Wilson et al. 2001). Most importantly, perhaps, AMPK-α1 activity was at least fourfold higher in second and third order branches of the pulmonary arterial tree when compared with systemic (mesenteric) arteries, which dilate rather than constrict in response to hypoxia. Once more this correlated with the distribution of the enzyme activities for the synthesis and metabolism of cADPR. Thus, the differential distribution of AMPK-α1 catalytic activity in arterial smooth muscle together with that of the enzyme activities for the synthesis and metabolism of cADPR could provide, via signal amplification, the degree of pulmonary selectivity required for HPV. Furthermore, the predominance of the α1β2γ1 heterotrimer in smooth muscle from second and third order branches of the pulmonary arterial tree suggests that this isozyme may play a prominent role in determining the response of pulmonary arteries to hypoxia.

Does hypoxia increase the AMP/ATP ratio and thereby activate AMPK-α1 activity in pulmonary arterial smooth muscle?

If inhibition of mitochondrial oxidative phosphorylation and consequent activation of AMPK were to mediate chemotransduction by hypoxia, we would expect a prerequisite increase in the AMP/ATP ratio on exposure of pulmonary arterial smooth muscle to hypoxia. This is exemplified by previous studies on the action of adenylate kinase, which have shown that any increase in the cellular ADP/ATP ratio is converted into a rise in the AMP/ATP ratio and consequent activation of AMPK. Consistent with this fact, we found that the AMP/ATP ratio in pulmonary arterial smooth muscle rose twofold under hypoxic conditions (16–21 mmHg; Evans et al. 2005). As one would expect (Hardie & Hawley, 2001; Hardie, 2003) the adenylate kinase reaction appeared close to equilibrium under normoxia and hypoxia, because the AMP/ATP ratio varied approximately as the square of the ADP/ATP ratio. Perhaps most significantly, the rise in the AMP/ATP ratio was associated with a concomitant, 2-fold increase in AMPK activity and phosphorylation of a classical AMPK substrate, acetyl CoA carboxylase. Notably, acute hypoxia was found to increase AMPK activity associated with the α1 catalytic subunit isoform to a greater extent than that associated with the α2 catalytic subunit isoform. This finding, when combined with the fact that AMPK-α1 activity was at least fourfold higher in pulmonary versus systemic arterial smooth muscle, provides further support for the view that the AMPK-α1 associated catalytic activity may play a prominent role in the response of pulmonary arterial smooth muscle cells to hypoxia.

AMPK, hypoxia and the regulation of the metabolic state in pulmonary versus systemic arteries

Previous investigations have shown that the energy state of pulmonary arterial smooth muscle is lower than that of systemic arteries, as defined by the phosphocreatine (PCr)/ATP and Pi/ATP ratios (Buescher et al. 1991; Leach et al. 1998, 2000). Upon exposure to hypoxia a fall in PCr levels and intracellular pH (pHi) was observed, whilst ATP levels were not altered significantly. These findings are significant because a lower energy state indicates that the resting AMP/ATP ratio may be higher in pulmonary arterial smooth muscle under normoxic conditions when compared to systemic arterial smooth muscle. The combined effect of variations in these metabolic markers could therefore be to provide a lower threshold for activation of AMPK upon exposure of pulmonary arteries to hypoxia. Consequently, a given level of hypoxia may facilitate AMPK-dependent signalling mechanisms to a greater extent in pulmonary when compared to systemic arterial smooth muscle, and the margin of differential signalling may be increased still further by the greater than fourfold higher levels of AMPK-α1 activity in pulmonary arterial smooth muscle.

It is also notable that upon exposure to hypoxia metabolic markers other than ATP (pHi, phosphocreatine) decline to a lesser degree in pulmonary arteries when compared to systemic arteries (Leach et al. 1998, 2000). This led to the suggestion that pulmonary arterial smooth muscle contains specific mechanisms that allow for better maintenance of the cellular energy state. Clearly, the ability of pulmonary arterial smooth muscle to maintain the energy state and phosphorylation potential could be conferred, at least in part, by the at least fourfold higher levels of AMPK-α1 activity relative to that found in systemic arterial smooth muscle, and by a lower threshold for AMPK activation by hypoxia. Indirect support for this view may be taken from previous investigations on pulmonary arterial smooth muscle, which have suggested that glucose uptake (governed by GLUT 4 in arterial smooth muscle; Banz et al. 1996) and glycolysis are accelerated by hypoxia, and that glycolysis plays a vital role not only in maintaining the energy status, but in maintaining pulmonary artery constriction by hypoxia (Ohe et al. 1986; Wiener & Sylvester, 1991; Zhao et al. 1996; Leach et al. 2000, 2001). This is because the primary role of AMPK in eukaryotic cells is to maintain ATP supply by activating catabolic processes and by inhibiting non-essential ATP consuming processes. Significantly, this is achieved in part by (1) accelerating glycolysis, (2) increasing glucose transport via the translocation of glucose transporters (including GLUT4) to the plasma membrane, and (3) inactivation of acetyl CoA carboxylase and consequent acceleration of β-oxidation of fatty acids by mitochondria (for further details see Hardie et al. 2006).

On the basis of the ability of pulmonary arterial smooth muscle to maintain the energy state and phosphorylation potential under hypoxia, it has also been argued that HPV is unlikely to be mediated by changes in metabolic variables (Buescher et al. 1991; Leach et al. 1998, 2000). It is vitally important therefore to recognize that AMPK is activated in response to a rise in the AMP/ATP ratio and that this can occur without a measurable fall in the cellular ATP levels. Furthermore, the maintenance of the phosphorylation potential will support the regulation of processes by AMPK-dependent phosphorylation of target proteins, such as those that contribute to stimulus–contraction coupling by hypoxia.

Does AMPK activation mobilize ryanodine-sensitive sarcoplasmic reticulum Ca2+ stores in pulmonary arterial smooth muscle?

Our investigations on the cellular distribution of the AMPK-α1 subunit isoform in isolated pulmonary arterial smooth muscle cells revealed that associated catalytic activity is likely to be targeted to sites throughout the cytoplasm, whilst being excluded from the nucleus and from the plasma membrane (Fig. 1A). Thus, a specific subunit isoform combination may target AMPK activity to SR compartments in order to support a role in the regulation of Ca2+ mobilization by AMPK in response to hypoxia. To determine whether or not AMPK did indeed mediate Ca2+ signalling by hypoxia, we employed the now classical method of studying AMPK activation. This requires the use of two drugs, phenformin and AICAR (5-aminoimidazole-4-carboxamide riboside), each of which activate AMPK via discrete mechanisms. Phenformin (a drug formerly used in the treatment of type 2 diabetes) inhibits Complex I of the mitochondrial respiratory chain (El-Mir et al. 2000; Owen et al. 2000) and thereby activates AMPK by increasing the cellular AMP/ATP ratio (Hawley et al. 2005). By contrast, AICAR is metabolized to yield the AMP mimetic ZMP (AICAR monophosphate) and thereby selectively activates AMPK without affecting the cellular AMP/ATP ratio (Corton et al. 1995; Owen et al. 2000). Despite their different modes of action, hypoxia (16–21 Torr), phenformin and AICAR activated AMPK and evoked an increase in intracellular Ca2+ concentration in isolated pulmonary arterial smooth muscle cells. In each case the increase in intracellular Ca2+ concentration was resistant to removal of extracellular Ca2+ (Dipp et al. 2001; Evans et al. 2005). However, prior block of SR Ca2+ release via RyRs by preincubation of cells with ryanodine and caffeine abolished the increase in intracellular Ca2+ concentration induced by each stimulus. Most significantly, SR Ca2+ release in response to AMPK activation was also abolished upon blocking the Ca2+ mobilizing messenger cADPR with 8-bromo-cADPR, a selective cADPR antagonist (Sethi et al. 1997; Walseth et al. 1997). Thus, AMPK activation triggers cADPR-dependent SR Ca2+ release via RyRs in isolated pulmonary arterial smooth muscle cells, as does hypoxia (Dipp & Evans, 2001; Dipp et al. 2001; Wilson et al. 2001). This finding was significant, because we had previously shown that the enzyme activities for the synthesis and metabolism of cADPR confer a degree of pulmonary selectivity, that cADPR accumulation in pulmonary arterial smooth muscle is augmented by hypoxia and that cADPR-dependent SR Ca2+ release is a prerequisite for HPV (Dipp & Evans, 2001; Wilson et al. 2001).

Figure 1. AMPK-dependent regulation of hypoxic pulmonary vasoconstriction.

Figure 1

A, from left to right: brightfield image of an isolated and fixed pulmonary arterial smooth muscle cell; z section showing immuno-labelling for the AMPK-α1 catalytic subunit isoform (green) and the DAPI labelled nucleus (blue); 3D reconstruction. B, schematic diagram describing the role of AMPK in HPV: hypoxia inhibits mitochondrial oxidative phosphorylation, which leads to an increase in the AMP/ATP ratio, AMPK activation, consequent cADPR-dependent SR Ca2+ release in pulmonary arterial smooth muscle and HPV.

We therefore sought to determine whether or not AMPK acts as a primary mediator of HPV. To this end we carried out a detailed comparison of the effects of hypoxia and AMPK activation by AICAR on isolated pulmonary arteries. Consistent with the time course of the maintained phase of pulmonary artery constriction by hypoxia (Dipp & Evans, 2001), AICAR induced a slow, sustained and reversible constriction of pulmonary artery rings (Evans et al. 2005). Removal of the pulmonary artery endothelium reduced the constriction in response to hypoxia and AICAR by approximately 30%. Furthermore, the endothelium-dependent component of constriction by AICAR and hypoxia was abolished upon removal of extracellular Ca2+, and therefore requires Ca2+ influx into the endothelium. In contrast, constriction mediated by mechanisms intrinsic to the smooth muscle was not so abolished. However, smooth muscle constriction was abolished by blocking the mobilization of SR Ca2+ stores via RyRs with ryanodine and caffeine and by blocking Ca2+ mobilization in response to cADPR with 8-bromo-cADPR. Thus, both AMPK activation and hypoxia mediate maintained constriction of pulmonary artery smooth muscle by cADPR-dependent mobilization of SR Ca2+ stores via RyRs (Fig. 1B).

It should be noted, however, that maintained smooth muscle constriction by AICAR and hypoxia, respectively, exhibited a partial dependence on transmembrane Ca2+ influx. In this respect it is of major significance that depletion of SR Ca2+ stores with caffeine and ryanodine or blockade of cADPR with 8-bromo-cADPR completely abolished the constriction of pulmonary arteries, with or without endothelium, by both AICAR and hypoxia. Thus, the partial dependence of smooth muscle constriction on extracellular Ca2+ must be determined by SR Ca2+ release-activated Ca2+ influx, as suggested by the investigations of others (Kang et al. 2003; Ng et al. 2005; Wang et al. 2005). However, it seems likely that this process of store-release activated Ca2+ influx is consequent to the mobilization of SR Ca2+ stores rather than being directly regulated by hypoxia (Kang et al. 2003; Ng et al. 2005) or AMPK activation per se (Evans et al. 2005).

The aforementioned findings provide strong support for our assertion that AMPK activation by hypoxia underpins HPV, and our most recent studies have provided further support for this view. Thus, the AMPK antagonist compound C (Zhou et al. 2001) inhibits the increase in acetyl CoA carboxylase phosphorylation in pulmonary arterial smooth muscle evoked by hypoxia and AICAR, and inhibits both HPV and pulmonary artery constriction in response to AICAR (A. M. Evans et al. unpublished results).

Does AMPK activation elicit voltage-gated Ca2+ influx into the carotid body type I cell and increased sensory afferent discharge from the carotid body?

In light of our observations on the role of AMPK in pulmonary artery constriction by hypoxia, we sought to determine the contribution of AMPK to chemotransduction by hypoxia in other O2-sensing cells, and focused our attention on the carotid body type I cell. These cells are also stimulated by hypoxia and, at least in part, via inhibition of mitochondrial oxidative phosphorylation. In this case, however, excitation is primarily mediated by voltage-gated Ca2+ influx, rather than by Ca2+ release from intracellular stores (Buckler & Vaughan-Jones, 1994). Surprisingly, immuno-fluorescence imaging revealed that the AMPK-α1 catalytic subunit isoform expressed in carotid body type I cells was almost entirely (∼75%) restricted to a volume within 1 μm of the plasma membrane (Fig. 2A; Evans et al. 2005). Thus, the spatial localization of AMPK-α1 in carotid body type I cells is consistent with it targeting plasma membrane delimited processes such as voltage-gated Ca2+ influx. This was contrary to our observations on pulmonary arterial smooth muscle cells, in which AMPK-α1 appeared absent from the plasma membrane and targeted to structures throughout the cytoplasm. One explanation for these conflicting observations is that AMPK-α1 forms a heterotrimer with different regulatory subunit isofoms (β and/or γ) in each of these cell types, and that the nature of the subunit combination determines the cellular compartment to which it is targeted. Consistent with this view, it has been suggested that the subunit isoform combination may determine the subcellular targeting of AMPK (Salt et al. 1998), and it has been proposed that post-translational modification of the regulatory β1 subunit isoform may target AMPK to the plasma membrane (Warden et al. 2001). The cellular distribution of AMPK may also be determined by the deployment of isozyme-specific anchoring proteins similar to the ‘A-kinase anchoring proteins’ (AKAPs) that underpin the cellular distribution of protein kinase A (Wang et al. 2006). One such candidate is plectin, a cytoskeleton linker protein that may play a role in the compartmentalization of intracellular signalling cascades (Janmey, 1998), which has been shown to bind γ1 regulatory subunits that are in combination with AMPK-α1 but not AMPK-α2 catalytic subunit isoforms, and in differentiated but not undifferentiated myotubes (Gregor et al. 2006). Thus, whilst the precise mechanisms remain to be determined it is clear that different AMPK isozymes may be targeted to discrete cellular compartments and in a cell-specific manner that may be tailored to the function of AMPK in a given cell type.

Figure 2. AMPK-dependent regulation of carotid body excitation by hypoxia.

Figure 2

A, from left to right: brightfield image of an isolated and fixed carotid body type I cell; z section showing immuno-labelling for the AMPK-α1 catalytic subunit isoform (green) tyrosine hydroxylase (red) and the DAPI labelled nucleus (blue); 3D reconstruction. B, schematic diagram describing the proposed mechanism of carotid body excitation by hypoxia: hypoxia inhibits mitochondrial oxidative phosphorylation in carotid body type I cells, which leads to an increase in the AMP/ATP ratio, AMPK activation, AMPK-dependent inhibition of K+ channels in the plasma membrane, voltage-gated Ca2+ influx, neurosecretion and a consequent increase in afferent fibre discharge.

Consistent with the effects of hypoxia and the distribution of AMPK-α1, AMPK activation by AICAR induced a reversible depolarization of the membrane potential of acutely isolated rat carotid body type I cells, which resulted from the selective inhibition of the O2-sensitive K+ currents in these cells, namely the TASK-like leak K+ current and BKCa current (Wyatt et al. 2006). Consequently, an increase in the intracellular Ca2+ concentration was evoked via Ni2+- and Cd2+-sensitive voltage-gated Ca2+ influx pathways, which ultimately led to an increase in sensory afferent discharge from the isolated carotid body (Evans et al. 2005; Wyatt et al. 2006). Thus, AMPK is likely to mediate the excitatory effects of hypoxia on isolated carotid body type I cells and on the carotid body in vitro (Fig. 2B). Once more this view gains strong support from our most recent investigations, which have shown that the AMPK antagonist compound C reverses the increase in intracellular Ca2+ concentration evoked by hypoxia and AICAR in isolated carotid body type I cells and the increase in sensory afferent discharge elicited from the isolated carotid body by each of these stimuli (A.M. Evans et al. unpublished results).

Could AMPK play a role in the development of hypoxic pulmonary hypertension and the effect of chronic intermittent hypoxia on hypoxic ventilatory responses?

Chronic and intermittent hypoxia precipitate (1) changes in hypoxic ventilatory responses, (2) remodelling of the carotid body, (3) pulmonary vascular remodelling associated with the development of hypoxic pulmonary hypertension, and (4) increased erythropoietin production and consequently increased red blood cell development. The transcription factor HIF-1 has been proposed to be pivotal to each of these pathophysiological alterations (Fandrey, 2004; Prabhakar & Jacono, 2005; Semenza, 2005). HIF-1 is a heterodimer comprising α and β subunits, the activity of which may be regulated by varying the expression of HIF-1α and thereby its association with the constitutively expressed β subunit. Therefore, It may be significant that AMPK-α2, which is preferentially targeted to the nucleus (Salt et al. 1998), but not the AMPK-α1 catalytic subunit isoform may increase HIF-1α expression (Hwang et al. 2004) and play a critical role in the regulation of associated transcriptional activity and target gene expression in response to prolonged hypoxia (Lee et al. 2003; Neurath et al. 2006). This raises the possibility that heterotrimeric subunit combinations containing AMPK-α1 may underpin the regulation of primary cell function by hypoxia as described above, whilst AMPK-α2 may play a prominent role in the paradigm of O2-dependent gene expression within O2-sensing cells in response to chronic and intermittent hypoxia.

Summary

In summary, previous studies have established that hypoxia promotes pulmonary artery constriction and carotid body sensory afferent discharge in part by inhibiting oxidative phosphorylation by mitochondria. Our findings now suggest that inhibition of mitochondrial oxidative phosphorylation by hypoxia leads to a rise in the cellular AMP/ATP ratio, consequent AMPK activation and the initiation of cell-specific Ca2+ signalling mechanisms in pulmonary arterial smooth muscle cells and carotid body type I cells. Thus, the characteristic response of each tissue type to AMPK activation precisely mirrors that seen in response to hypoxia: (1) constriction of pulmonary arteries by cADPR-dependent SR Ca2+ release in the smooth muscle cells, with a secondary component of constriction driven by the pulmonary arterial endothelial cells, and (2) inhibition of TASK-like and BKCa K+ currents, leading to depolarization, voltage-gated Ca2+ influx into carotid body type I cells and a consequent increase in afferent fibre discharge. We therefore proposed that in addition to maintaining the cellular energy state, AMPK acts as the primary metabolic sensor and effector of Ca2+ signalling by hypoxia in O2-sensing cells. This view unites for the first time the mitochondrial and Ca2+ signalling hypotheses for chemotransduction by hypoxia. Our data also suggest that high levels of AMPK-α1 expression, together with a reliance on mitochondrial oxidative phosphorylation for ATP production, may explain why O2-sensing cells are particularly sensitive to hypoxia. It is also possible that a lower resting energy state in O2-sensing cells relative to other cell types may alter the metabolic setting in such a way that may increase the sensitivity of these cells to hypoxia by lowering the threshold for activation of AMPK.

Given the ubiquitous expression of AMPK in eukaryotes, it is intriguing to speculate on the possibility that the targeting of AMPK-dependent Ca2+ signalling mechanisms to specialized cells that function to monitor O2 supply may have been a relatively early evolutionary development. This may be revealed by future studies on the role of AMPK as a metabolic sensor and effector of chemotransduction by hypoxia in O2-sensing cells and tissues from primitive species; for example, systemic arteries of cyclostomes, arteries that feed the skin of amphibia and neuroepithelial cells from the gills of teleosts. If this proves to be the case, it would suggest that AMPK-dependent signalling mechanisms have likely been adapted to serve a variety of cell- and tissue-specific functions for which they are fit for purpose.

Acknowledgments

I would like to thank Professor Chris Peers, Professor D. Grahame Hardie and Dr Prem Kumar for their enthusiastic support and encouragement throughout our collaborative studies on the role of AMPK in chemotransduction by hypoxia.

References

  1. Aley PK, Murray HJ, Boyle JP, Pearson HA, Peers C. Hypoxia stimulates Ca2+ release from intracellular stores in astrocytes via cyclic ADP ribose-mediated activation of ryanodine receptors. Cell Calcium. 2006;39:95–100. doi: 10.1016/j.ceca.2005.09.009. [DOI] [PubMed] [Google Scholar]
  2. Aley PK, Porter KE, Boyle JP, Kemp PJ, Peers C. Hypoxic modulation of Ca2+ signaling in human venous endothelial cells. Multiple roles for reactive oxygen species. J Biol Chem. 2005;280:13349–13354. doi: 10.1074/jbc.M413674200. [DOI] [PubMed] [Google Scholar]
  3. Anderson KA, Means RL, Huang QH, Kemp BE, Goldstein EG, Selbert MA, Edelman AM, Fremeau RT, Means AR. Components of a calmodulin-dependent protein kinase cascade. Molecular cloning, functional characterization and cellular localization of Ca2+/calmodulin-dependent protein kinase kinase β. J Biol Chem. 1998;273:31880–31889. doi: 10.1074/jbc.273.48.31880. [DOI] [PubMed] [Google Scholar]
  4. Archer SL, Huang J, Henry T, Peterson D, Weir EK. A redox-based O2 sensor in rat pulmonary vasculature. Circ Res. 1993;73:1100–1112. doi: 10.1161/01.res.73.6.1100. [DOI] [PubMed] [Google Scholar]
  5. Archer SL, Will JA, Weir EK. Redox status in the control of pulmonary vascular tone. Herz. 1986;11:127–141. [PubMed] [Google Scholar]
  6. Banz WJ, Abel MA, Zemel MB. Insulin regulation of vascular smooth muscle glucose transport in insulin-sensitive and resistant rats. Horm Metab Res. 1996;28:271–275. doi: 10.1055/s-2007-979790. [DOI] [PubMed] [Google Scholar]
  7. Buckler KJ. A novel oxygen-sensitive potassium current in rat carotid body type I cells. J Physiol. 1997;498:649–662. doi: 10.1113/jphysiol.1997.sp021890. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Buckler KJ. Background leak K+-currents and oxygen sensing in carotid body type 1 cells. Respir Physiol. 1999;115:179–187. doi: 10.1016/s0034-5687(99)00015-8. [DOI] [PubMed] [Google Scholar]
  9. Buckler KJ, Vaughan-Jones RD. Effects of hypercapnia on membrane potential and intracellular calcium in rat carotid body type I cells. J Physiol. 1994;478:157–171. doi: 10.1113/jphysiol.1994.sp020239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Buckler KJ, Williams BA, Honore E. An oxygen-, acid- and anaesthetic-sensitive TASK-like background potassium channel in rat arterial chemoreceptor cells. J Physiol. 2000;525:135–142. doi: 10.1111/j.1469-7793.2000.00135.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Buescher PC, Pearse DB, Pillai RP, Litt MC, Mitchell MC, Sylvester JT. Energy state and vasomotor tone in hypoxic pig lungs. J Appl Physiol. 1991;70:1874–1881. doi: 10.1152/jappl.1991.70.4.1874. [DOI] [PubMed] [Google Scholar]
  12. Chen J, Gomez-Nino A, Gonzalez C, Dinger B, Fidone S. Stimulus-specific mobilization of dopamine and norepinephrine stores in cat carotid body. J Auton Nerv Syst. 1997;67:109–113. doi: 10.1016/s0165-1838(97)00095-7. [DOI] [PubMed] [Google Scholar]
  13. Cheung PC, Salt IP, Davies SP, Hardie DG, Carling D. Characterization of AMP-activated protein kinase γ-subunit isoforms and their role in AMP binding. Biochem J. 2000;346:659–669. [PMC free article] [PubMed] [Google Scholar]
  14. Corton JM, Gillespie JG, Hawley SA, Hardie DG. 5-Aminoimidazole-4-carboxamide ribonucleoside: a specific method for activating AMP-activated protein kinase in intact cells? Eur J Biochem. 1995;229:558–565. doi: 10.1111/j.1432-1033.1995.tb20498.x. [DOI] [PubMed] [Google Scholar]
  15. Demiryurek AT, Wadsworth RM, Kane KA. Pharmacological evidence for the role of mediators in hypoxia-induced vasoconstriction in sheep isolated arteries. Eur J Pharmacol. 1991;203:1–8. doi: 10.1016/0014-2999(91)90783-m. [DOI] [PubMed] [Google Scholar]
  16. Dipp M, Evans AM. Cyclic ADP-ribose is the primary trigger for hypoxic pulmonary vasoconstriction in the rat lung in situ. Circ Res. 2001;89:77–83. doi: 10.1161/hh1301.093616. [DOI] [PubMed] [Google Scholar]
  17. Dipp M, Nye PCG, Evans AM. Hypoxic release of calcium from the sarcoplasmic reticulum of pulmonary artery smooth muscle. Am J Physiol. 2001;281:L318–L325. doi: 10.1152/ajplung.2001.281.2.L318. [DOI] [PubMed] [Google Scholar]
  18. Duchen MR, Biscoe TJ. Mitochondrial function in type I cells isolated from rabbit arterial chemoreceptors. J Physiol. 1992a;450:13–31. doi: 10.1113/jphysiol.1992.sp019114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Duchen MR, Biscoe TJ. Relative mitochondrial membrane potential and [Ca2+]i in type I cells isolated from the rabbit carotid body. J Physiol. 1992b;450:33–61. doi: 10.1113/jphysiol.1992.sp019115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. El-Mir MY, Nogueira V, Fontaine E, Averet N, Rigoulet M, Leverve X. Dimethylbiguanide inhibits cell respiration via an indirect effect targeted on the respiratory chain complex I. J Biol Chem. 2000;275:223–228. doi: 10.1074/jbc.275.1.223. [DOI] [PubMed] [Google Scholar]
  21. Evans AM. Hypoxia, cell metabolism, and cADPR accumulation. In: Yuan X-J, editor. Hypoxic Pulmonary Vasoconstriction: Cellular and Molecular Mechanisms. Boston: Kluwer Academic Publishers; 2004. pp. 313–338. [Google Scholar]
  22. Evans AM, Hardie DG, Galione A, Peers C, Kumar P, Wyatt CN. AMP-activated protein kinase couples mitochondrial inhibition by hypoxia to cell-specific Ca+ signalling mechanisms in oxygen-sensing cells. Novartis Found Symp. 2006a;272:234–252. discussion 252-8, 274-9. [PubMed] [Google Scholar]
  23. Evans AM, Mustard KJW, Wyatt CN, Dipp M, Kinnear NP, Hardie DG. Does AMP-activated protein kinase couple inhibition of mitochondrial oxidative phosphorylation by hypoxia to pulmonary artery constriction. Adv Exp Med Biol. 2006b;580:147–154. doi: 10.1007/0-387-31311-7_22. discussion 351-149. [DOI] [PubMed] [Google Scholar]
  24. Evans AM, Mustard KJ, Wyatt CN, Peers C, Dipp M, Kumar P, Kinnear NP, Hardie DG. Does AMP-activated protein kinase couple inhibition of mitochondrial oxidative phosphorylation by hypoxia to calcium signaling in O2-sensing cells? J Biol Chem. 2005;280:41504–41511. doi: 10.1074/jbc.M510040200. [DOI] [PubMed] [Google Scholar]
  25. Fandrey J. Oxygen-dependent and tissue-specific regulation of erythropoietin gene expression. Am J Physiol Regul Integr Comp Physiol. 2004;286:R977–R988. doi: 10.1152/ajpregu.00577.2003. [DOI] [PubMed] [Google Scholar]
  26. Fidone SJ, Gonzalez C, Dinger BG, Hanson GR. Mechanisms of chemotransmission in the mammalian carotid body. Prog Brain Res. 1988;74:169–179. doi: 10.1016/s0079-6123(08)63011-3. [DOI] [PubMed] [Google Scholar]
  27. Fishman AP. Hypoxia on the pulmonary circulation. How and where it acts. Circ Res. 1976;38:221–231. doi: 10.1161/01.res.38.4.221. [DOI] [PubMed] [Google Scholar]
  28. Gaine SP, Hales MA, Flavahan NA. Hypoxic pulmonary endothelial cells release a diffusible contractile factor distinct from endothelin. Am J Physiol. 1998;274:L657–L664. doi: 10.1152/ajplung.1998.274.4.L657. [DOI] [PubMed] [Google Scholar]
  29. Galione A, Lee HC, Busa WB. Ca2+-induced Ca2+ release in sea urchin egg homogenates: modulation by cyclic ADP-ribose. Science. 1991;253:1143–1146. doi: 10.1126/science.1909457. [DOI] [PubMed] [Google Scholar]
  30. Ganfornina MD, Lopez-Barneo J. Single K+ channels in membrane patches of arterial chemoreceptor cells are modulated by O2 tension. Proc Natl Acad Sci U S A. 1991;88:2927–2930. doi: 10.1073/pnas.88.7.2927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Ganfornina MD, Lopez-Barneo J. Gating of O2-sensitive K+ channels of arterial chemoreceptor cells and kinetic modifications induced by low Po2. J General Physiol. 1992;100:427–455. doi: 10.1085/jgp.100.3.427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Gnaiger E, Lassnig B, Kuznetsov A, Rieger G, Margreiter R. Mitochondrial oxygen affinity, respiratory flux control and excess capacity of cytochrome c oxidase. J Exp Biol. 1998;201:1129–1139. doi: 10.1242/jeb.201.8.1129. [DOI] [PubMed] [Google Scholar]
  33. 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]
  34. Gonzalez C, Sanz-Alfayate G, Agapito MT, Gomez-Nino A, Rocher A, Obeso A. Significance of ROS in oxygen sensing in cell systems with sensitivity to physiological hypoxia. Respir Physiol Neurobiol. 2002;132:17–41. doi: 10.1016/s1569-9048(02)00047-2. [DOI] [PubMed] [Google Scholar]
  35. Gregor M, Zeold A, Oehler S, Marobela KA, Fuchs P, Weigel G, Hardie DG, Wiche G. Plectin scaffolds recruit energy-controlling AMP-activated protein kinase (AMPK) in differentiated myofibres. J Cell Sci. 2006;119:1864–1875. doi: 10.1242/jcs.02891. [DOI] [PubMed] [Google Scholar]
  36. Hardie DG. Minireview: The AMP-activated protein kinase cascade: the key sensor of cellular energy status. Endocrinology. 2003;144:5179–5183. doi: 10.1210/en.2003-0982. [DOI] [PubMed] [Google Scholar]
  37. Hardie DG. New roles for the LKB1→AMPK pathway. Curr Opin Cell Biol. 2005;17:167–173. doi: 10.1016/j.ceb.2005.01.006. [DOI] [PubMed] [Google Scholar]
  38. Hardie DG, Hawley SA. AMP-activated protein kinase: the energy charge hypothesis revisited. Bioessays. 2001;23:1112–1119. doi: 10.1002/bies.10009. [DOI] [PubMed] [Google Scholar]
  39. Hardie DG, Hawley SA, Scott JW. AMP-activated protein kinase – development of the energy sensor concept. J Physiol. 2006;574:7–15. doi: 10.1113/jphysiol.2006.108944. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Hawley SA, Boudeau J, Reid JL, Mustard KJ, Udd L, Makela TP, Alessi DR, Hardie DG. Complexes between the LKB1 tumor suppressor, STRADα/β and MO25α/β are upstream kinases in the AMP-activated protein kinase cascade. J Biol. 2003;2:28. doi: 10.1186/1475-4924-2-28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Hawley SA, Pan DA, Mustard KJ, Ross L, Bain J, Edelman AM, Frenguelli BG, Hardie DG. Calmodulin-dependent protein kinase kinase-β is an alternative upstream kinase for AMP-activated protein kinase. Cell Metab. 2005;2:9–19. doi: 10.1016/j.cmet.2005.05.009. [DOI] [PubMed] [Google Scholar]
  42. Hawley SA, Selbert MA, Goldstein EG, Edelman AM, Carling D, Hardie DG. 5′-AMP activates the AMP-activated protein kinase cascade, and Ca2+/calmodulin activates the calmodulin-dependent protein kinase I cascade, via three independent mechanisms. J Biol Chem. 1995;270:27186–27191. doi: 10.1074/jbc.270.45.27186. [DOI] [PubMed] [Google Scholar]
  43. He L, Dinger B, Sanders K, Hoidal J, Obeso A, Stensaas L, Fidone S, Gonzalez C. Effect of p47phox gene deletion on ROS production and oxygen sensing in mouse carotid body chemoreceptor cells. Am J Physiol Lung Cell Mol Physiol. 2005;289:L916–L924. doi: 10.1152/ajplung.00015.2005. [DOI] [PubMed] [Google Scholar]
  44. Heymans C, Boukaert JJ, Dautrebande L. Sinus carotidien et reflexes respiratoires. II. Influences respiratoires reflexes de l'acidose, de l'acolose, de l'anhydride carbonique, de l'ion hydrogene et de l'anoxemie: sinus carotidiens et echanges respiratoires dans les poumans et au dela de poumons. Arch Int Pharmacodyn Ther. 1930;39:400–408. [Google Scholar]
  45. Hurley RL, Anderson KA, Franzone JM, Kemp BE, Means AR, Witters LA. The Ca2+/calmoldulin-dependent protein kinase kinases are AMP-activated protein kinase kinases. J Biol Chem. 2005;280:29060–29066. doi: 10.1074/jbc.M503824200. [DOI] [PubMed] [Google Scholar]
  46. Hwang JT, Lee M, Jung SN, Lee HJ, Kang I, Kim SS, Ha J. AMP-activated protein kinase activity is required for vanadate-induced hypoxia-inducible factor 1α expression in DU145 cells. Carcinogenesis. 2004;25:2497–2507. doi: 10.1093/carcin/bgh253. [DOI] [PubMed] [Google Scholar]
  47. Jabr RI, Toland H, Gelband CH, Wang XX, Hume JR. Prominent role of intracellular Ca2+ release in hypoxic pulmonary vasoconstriction of canine pulmonary artery. Br J Pharmacol. 1997;122:21–30. doi: 10.1038/sj.bjp.0701326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Janmey PA. The cytoskeleton and cell signaling: component localization and mechanical coupling. Physiol Rev. 1998;78:763–781. doi: 10.1152/physrev.1998.78.3.763. [DOI] [PubMed] [Google Scholar]
  49. Jonz MG, Fearon IM, Nurse CA. Neuroepithelial oxygen chemoreceptors of the zebrafish gill. J Physiol. 2004;560:737–752. doi: 10.1113/jphysiol.2004.069294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Kang TM, Park MK, Uhm DY. Effects of hypoxia and mitochondrial inhibition on the capacitative calcium entry in rabbit pulmonary arterial smooth muscle cells. Life Sci. 2003;72:1467–1479. doi: 10.1016/s0024-3205(02)02441-4. [DOI] [PubMed] [Google Scholar]
  51. Kato M, Staub NC. Response of small pulmonary arteries to unilobar hypoxia and hypercania. Circ Res. 1966;19:426–440. doi: 10.1161/01.res.19.2.426. [DOI] [PubMed] [Google Scholar]
  52. Kemp BE. Bateman domains and adenosine derivatives form a binding contract. J Clin Invest. 2004;113:182–184. doi: 10.1172/JCI20846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Killilea DW, Hester R, Balczon R, Babal P, Gillespie MN. Free radical production in hypoxic pulmonary artery smooth muscle cells. Am J Physiol Lung Cell Mol Physiol. 2000;279:L408–L412. doi: 10.1152/ajplung.2000.279.2.L408. [DOI] [PubMed] [Google Scholar]
  54. Kovitz KL, Aleskowitch TD, Sylvester JT, Flavahan NA. Endothelium-derived contracting and relaxing factors contribute to hypoxic responses of pulmonary arteries. Am J Physiol. 1993;265:H1139–H1148. doi: 10.1152/ajpheart.1993.265.4.H1139. [DOI] [PubMed] [Google Scholar]
  55. Leach RM, Hill HM, Snetkov VA, Robertson TP, Ward JP. Divergent roles of glycolysis and the mitochondrial electron transport chain in hypoxic pulmonary vasoconstriction of the rat: identity of the hypoxic sensor. J Physiol. 2001;536:211–224. doi: 10.1111/j.1469-7793.2001.00211.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Leach RM, Sheehan DW, Chacko VP, Sylvester JT. Effects of hypoxia on energy state and pH in resting pulmonary and femoral arterial smooth muscles. Am J Physiol. 1998;275:L1051–L1060. doi: 10.1152/ajplung.1998.275.6.L1051. [DOI] [PubMed] [Google Scholar]
  57. Leach RM, Sheehan DW, Chacko VP, Sylvester JT. Energy state, pH, and vasomotor tone during hypoxia in precontracted pulmonary and femoral arteries. Am J Physiol. 2000;278:L294–L304. doi: 10.1152/ajplung.2000.278.2.L294. [DOI] [PubMed] [Google Scholar]
  58. Lee M, Hwang JT, Lee HJ, Jung SN, Kang I, Chi SG, Kim SS, Ha J. AMP-activated protein kinase activity is critical for hypoxia-inducible factor-1 transcriptional activity and its target gene expression under hypoxic conditions in DU145 cells. J Biol Chem. 2003;278:39653–39661. doi: 10.1074/jbc.M306104200. [DOI] [PubMed] [Google Scholar]
  59. Lee HC, Walseth TF, Bratt GT, Hayes RN, Clapper DL. Structural determination of a cyclic metabolite of NAD+ with intracellular Ca2+-mobilizing activity. J Biol Chem. 1989;264:1608–1615. [PubMed] [Google Scholar]
  60. Lopez-Barneo J, Lopez-Lopez JR, Urena J, Gonzalez C. Chemotransduction in the carotid body: K+ current modulated by PO2 in type I chemoreceptor cells. Science. 1988;241:580–582. doi: 10.1126/science.2456613. [DOI] [PubMed] [Google Scholar]
  61. Lopez-Lopez JR, Gonzalez C. Time course of K+ current inhibition by low oxygen in chemoreceptor cells of adult rabbit carotid body. Effects of carbon monoxide. FEBS Lett. 1992;299:251–254. doi: 10.1016/0014-5793(92)80126-2. [DOI] [PubMed] [Google Scholar]
  62. Lopez-Lopez JR, Perez-Garcia MT, Sanz-Alfayate G, Obeso A, Gonzalez C. Functional identification of Kvα subunits contributing to the O2-sensitive K+ current in rabbit carotid body chemoreceptor cells. Adv Exp Med Biol. 2003;536:33–39. doi: 10.1007/978-1-4419-9280-2_4. [DOI] [PubMed] [Google Scholar]
  63. Malvin GM, Walker BR. Sites and ionic mechanisms of hypoxic vasoconstriction in frog skin. Am J Physiol Regul Integr Comp Physiol. 2001;280:R1308–R1314. doi: 10.1152/ajpregu.2001.280.5.R1308. [DOI] [PubMed] [Google Scholar]
  64. Marsin AS, Bertrand L, Rider MH, Deprez J, Beauloye C, Vincent MF, Van den Berghe G, Carling D, Hue L. Phosphorylation and activation of heart PFK-2 by AMPK has a role in the stimulation of glycolysis during ischaemia. Curr Biol. 2000;10:1247–1255. doi: 10.1016/s0960-9822(00)00742-9. [DOI] [PubMed] [Google Scholar]
  65. Michelakis ED, Thebaud B, Weir EK, Archer SL. Hypoxic pulmonary vasoconstriction: redox regulation of O2-sensitive K+ channels by a mitochondrial O2-sensor in resistance artery smooth muscle cells. J Mol Cell Cardiol. 2004;37:1119–1136. doi: 10.1016/j.yjmcc.2004.09.007. [DOI] [PubMed] [Google Scholar]
  66. Mills E, Jobsis FF. Mitochondrial respiratory chain of carotid body and chemoreceptor response to changes in oxygen tension. J Neurophysiol. 1972;35:405–428. doi: 10.1152/jn.1972.35.4.405. [DOI] [PubMed] [Google Scholar]
  67. Morio Y, McMurtry IF. Ca2+ release from ryanodine-sensitive store contributes to mechanism of hypoxic vasoconstriction in rat lungs. J Appl Physiol. 2002;92:527–534. doi: 10.1152/jappl.2002.92.2.527. [DOI] [PubMed] [Google Scholar]
  68. Neurath KM, Keough MP, Mikkelsen T, Claffey KP. AMP-dependent protein kinase α2 isoform promotes hypoxia-induced VEGF expression in human glioblastoma. Glia. 2006;53:733–743. doi: 10.1002/glia.20326. [DOI] [PubMed] [Google Scholar]
  69. Ng LC, Wilson SM, Hume JR. Mobilization of sarcoplasmic reticulum stores by hypoxia leads to consequent activation of capacitative Ca2+ entry in isolated canine pulmonary arterial smooth muscle cells. J Physiol. 2005;563:409–419. doi: 10.1113/jphysiol.2004.078311. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Ohe M, Mimata T, Haneda T, Takishima T. Time course of pulmonary vasoconstriction with repeated hypoxia and glucose depletion. Respir Physiol. 1986;63:177–186. doi: 10.1016/0034-5687(86)90112-x. [DOI] [PubMed] [Google Scholar]
  71. Olson KR, Russell MJ, Forster ME. Hypoxic vasoconstriction of cyclostome systemic vessels: the antecedent of hypoxic pulmonary vasoconstriction? Am J Physiol Regul Integr Comp Physiol. 2001;280:R198–R206. doi: 10.1152/ajpregu.2001.280.1.R198. [DOI] [PubMed] [Google Scholar]
  72. Owen MR, Doran E, Halestrap AP. Evidence that metformin exerts its anti-diabetic effects through inhibition of complex 1 of the mitochondrial respiratory chain. Biochem J. 2000;348:607–614. [PMC free article] [PubMed] [Google Scholar]
  73. Peers C. Hypoxic suppression of K+ currents in type I carotid body cells: selective effect on the Ca2+-activated K+ current. Neurosci Lett. 1990;119:253–256. doi: 10.1016/0304-3940(90)90846-2. [DOI] [PubMed] [Google Scholar]
  74. Perez-Garcia MT, Colinas O, Miguel-Velado E, Moreno-Dominguez A, Lopez-Lopez JR. Characterization of the Kv channels of mouse carotid body chemoreceptor cells and their role in oxygen sensing. J Physiol. 2004;557:457–471. doi: 10.1113/jphysiol.2004.062281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Prabhakar NR. Endogenous carbon monoxide in control of respiration. Respir Physiol. 1998;114:57–64. doi: 10.1016/s0034-5687(98)00072-3. [DOI] [PubMed] [Google Scholar]
  76. 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]
  77. Prabhakar NR, Kumar GK. Oxidative stress in the systemic and cellular responses to intermittent hypoxia. Biol Chem. 2004;385:217–221. doi: 10.1515/BC.2004.015. [DOI] [PubMed] [Google Scholar]
  78. Prabhakar NR, Overholt JL. Cellular mechanisms of oxygen sensing at the carotid body: heme proteins and ion channels. Respir Physiol. 2000;122:209–221. doi: 10.1016/s0034-5687(00)00160-2. [DOI] [PubMed] [Google Scholar]
  79. Robertson TP, Aaronson PI, Ward JPT. Hypoxic vasoconstriction and intracellular Ca2+ in pulmonary arteries: evidence for PKC-independent Ca2+ sensitization. Am J Physiol. 1995;268:H301–H307. doi: 10.1152/ajpheart.1995.268.1.H301. [DOI] [PubMed] [Google Scholar]
  80. Robertson TP, Ward JPT, Aaronson PI. Hypoxia induces the release of a pulmonary-selective, Ca2+-sensitising, vasoconstrictor from the perfused rat lung. Cardiovasc Res. 2001;50:145–150. doi: 10.1016/s0008-6363(01)00192-4. [DOI] [PubMed] [Google Scholar]
  81. Rounds S, McMurtry IF. Inhibitors of oxidative ATP production cause transient vasoconstriction and block subsequent pressor responses in rat lungs. Circ Res. 1981;48:393–400. doi: 10.1161/01.res.48.3.393. [DOI] [PubMed] [Google Scholar]
  82. Roy CS, Sherrington CS. The regulation of blood supply of the brain. J Physiol. 1890;11:85–108. doi: 10.1113/jphysiol.1890.sp000321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Rusinko N, Lee HC. Widespread occurrence in animal tissues of an enzyme catalysing the conversion of NAD+ into a cyclic metabolite with intracellular Ca2+ mobilising activity. J Biol Chem. 1989;264:11725–11731. [PubMed] [Google Scholar]
  84. Sakamoto K, McCarthy A, Smith D, Green KA, Hardie DG, Ashworth A, Alessi DR. Deficiency of LKB1 in skeletal muscle prevents AMPK activation and glucose uptake during contraction. EMBO J. 2005;24:1810–1820. doi: 10.1038/sj.emboj.7600667. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Sakamoto K, Zarrinpashneh E, Budas GR, Pouleur AC, Dutta A, Prescott AR, Ashworth A, Jovanovic A, Alessi DR, Bertrand L. Deficiency of LKB1 in heart prevents ischemia-mediated activation of AMPKα2 but not AMPKα1. Am J Physiol. 2006;290:E780–788. doi: 10.1152/ajpendo.00443.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Salt IP, Celler JW, Hawley SA, Prescott A, Woods A, Carling D, Hardie DG. AMP-activated protein kinase – greater AMP dependence, and preferential nuclear localization, of complexes containing the α2 isoform. Biochem J. 1998;334:177–187. doi: 10.1042/bj3340177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Salvaterra CG, Goldman WF. Acute hypoxia increases cytosolic calcium in cultured pulmonary arterial myocytes. Am J Physiol. 1993;1993:L323–L328. doi: 10.1152/ajplung.1993.264.3.L323. [DOI] [PubMed] [Google Scholar]
  88. Sanchez D, Lopez-Lopez JR, Perez-Garcia MT, Sanz-Alfayate G, Obeso A, Ganfornina MD, Gonzalez C. Molecular identification of Kvα subunits that contribute to the oxygen-sensitive K+ current of chemoreceptor cells of the rabbit carotid body. J Physiol. 2002;542:369–382. doi: 10.1113/jphysiol.2002.018382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Scott JW, Norman DG, Hawley SA, Kontogiannis L, Hardie DG. Protein kinase substrate recognition studied using the recombinant catalytic domain of AMP-activated protein kinase and a model substrate. J Mol Biol. 2002;317:309–323. doi: 10.1006/jmbi.2001.5316. [DOI] [PubMed] [Google Scholar]
  90. Semenza GL. Involvement of hypoxia-inducible factor 1 in pulmonary pathophysiology. Chest. 2005;128:592S–594S. doi: 10.1378/chest.128.6_suppl.592S. [DOI] [PubMed] [Google Scholar]
  91. Sethi JK, Empson RM, Bailey VC, Potter BVL, Galione A. 7-Deaza-8-bromo-cyclic ADP-ribose, the first membrane-permeant, hydrolysis-resistant cyclic ADP-ribose antagonist. J Biol Chem. 1997;272:16358–16363. doi: 10.1074/jbc.272.26.16358. [DOI] [PubMed] [Google Scholar]
  92. Shaw RJ, Kosmatka M, Bardeesy N, Hurley RL, Witters LA, DePinho RA, Cantley LC. The tumor suppressor LKB1 kinase directly activates AMP-activated kinase and regulates apoptosis in response to energy stress. Proc Natl Acad Sci U S A. 2004;101:3329–3335. doi: 10.1073/pnas.0308061100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Shaw RJ, Lamia KA, Vasquez D, Koo SH, Bardeesy N, Depinho RA, Montminy M, Cantley LC. The kinase LKB1 mediates glucose homeostasis in liver and therapeutic effects of metformin. Science. 2005;310:1642–1646. doi: 10.1126/science.1120781. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Shigemori K, Ishizaki T, Matsukawa S, Sakai A, Nakai T, Miyabo S. Adenine nucleotides via activation of ATP-sensitive K+ channels modulate hypoxic response in rat pulmonary artery. Am J Physiol. 1996;270:L803–L809. doi: 10.1152/ajplung.1996.270.5.L803. [DOI] [PubMed] [Google Scholar]
  95. Stea A, Nurse CA. Whole-cell and perforated-patch recordings from O2-sensitive rat carotid body cells grown in short- and long-term culture. Pflugers Arch. 1991;418:93–101. doi: 10.1007/BF00370457. [DOI] [PubMed] [Google Scholar]
  96. Voelkel NF. Mechanisms of hypoxic pulmonary vasoconstriction. Am Rev Respir Dis. 1986;133:1186–1195. doi: 10.1164/arrd.1986.133.6.1186. [DOI] [PubMed] [Google Scholar]
  97. von Euler US, Liljestrand G. Observations on the pulmonary arterial blood pressure in the cat. Acta Physiol Scand. 1946;12:301–320. [Google Scholar]
  98. Walseth TF, Aarhus R, Gurnack ME, Wong L, Breitinger HG, Gee KR, Lee HC. Preparation of cyclic ADP-ribose antagonists and caged cyclic ADP-ribose. Meth Enzymol. 1997;280:294–305. doi: 10.1016/s0076-6879(97)80121-8. [DOI] [PubMed] [Google Scholar]
  99. Walseth TF, Aarhus R, Zeleznicker RJ, Lee HC. Determination of endogenous levels of cyclic ADP-ribose in rat tissue. Biochim Biophys Acta. 1991;1094:113–120. doi: 10.1016/0167-4889(91)90032-s. [DOI] [PubMed] [Google Scholar]
  100. Wang J, Shimoda LA, Weigand L, Wang W, Sun D, Sylvester JT. Acute hypoxia increases intracellular [Ca2+] in pulmonary arterial smooth muscle by enhancing capacitative Ca2+ entry. Am J Physiol Lung Cell Mol Physiol. 2005;288:L1059–L1069. doi: 10.1152/ajplung.00448.2004. [DOI] [PubMed] [Google Scholar]
  101. Wang HY, Tao J, Shumay E, Malbon CC. G-protein-coupled receptor-associated A-kinase anchoring proteins: AKAP79 and AKAP250 (gravin) Eur J Cell Biol. 2006 doi: 10.1016/j.ejcb.2005.12.003. in press. [DOI] [PubMed] [Google Scholar]
  102. Warden SM, Richardson C, O'Donnell J, Jr, Stapleton D, Kemp BE, Witters LA. Post-translational modifications of the β-1 subunit of AMP-activated protein kinase affect enzyme activity and cellular localization. Biochem J. 2001;354:275–283. doi: 10.1042/0264-6021:3540275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Waypa GB, Chandel NS, Schumacker PT. Model for hypoxic pulmonary vasoconstriction involving mitochondrial oxygen sensing. Circ Res. 2001;88:1259–1266. doi: 10.1161/hh1201.091960. [DOI] [PubMed] [Google Scholar]
  104. Wiener CM, Sylvester JT. Effects of glucose on hypoxic vasoconstriction in isolated ferret lungs. J Appl Physiol. 1991;70:439–446. doi: 10.1152/jappl.1991.70.1.439. [DOI] [PubMed] [Google Scholar]
  105. Williams SE, Wootton P, Mason HS, Bould J, Iles DE, Riccardi D, Peers C, Kemp PJ. Hemoxygenase-2 is an oxygen sensor for a calcium-sensitive potassium channel. Science. 2004;306:2093–2097. doi: 10.1126/science.1105010. [DOI] [PubMed] [Google Scholar]
  106. Wilson HL, Dipp M, Thomas JM, Lad C, Galione A, Evans AM. ADP-ribosyl cyclase and cyclic ADP-ribose hydrolase act as a redox sensor: a primary role for cADPR in hypoxic pulmonary vasoconstriction. J Biol Chem. 2001;276:11180–11188. doi: 10.1074/jbc.M004849200. [DOI] [PubMed] [Google Scholar]
  107. Winder WW, Hardie DG. Inactivation of acetyl-CoA carboxylase and activation of AMP-activated protein kinase in muscle during exercise. Am J Physiol. 1996;270:E299–E304. doi: 10.1152/ajpendo.1996.270.2.E299. [DOI] [PubMed] [Google Scholar]
  108. Woods A, Dickerson K, Heath R, Hong SP, Momcilovic M, Johnstone SR, Carlson M, Carling D. Ca2+/calmodulin-dependent protein kinase kinase-β acts upstream of AMP-activated protein kinase in mammalian cells. Cell Metab. 2005;2:21–33. doi: 10.1016/j.cmet.2005.06.005. [DOI] [PubMed] [Google Scholar]
  109. Woods A, Johnstone SR, Dickerson K, Leiper FC, Fryer LG, Neumann D, Schlattner U, Wallimann T, Carlson M, Carling D. LKB1 is the upstream kinase in the AMP-activated protein kinase cascade. Curr Biol. 2003;13:2004–2008. doi: 10.1016/j.cub.2003.10.031. [DOI] [PubMed] [Google Scholar]
  110. Wyatt CN, Buckler KJ. The effect of mitochondrial inhibitors on membrane currents in isolated neonatal rat carotid body type I cells. J Physiol. 2004;556:175–191. doi: 10.1113/jphysiol.2003.058131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Wyatt CN, Peers C. Modulation of ionic currents in isolated type I cells of the neonatal rat carotid body by p-chloromercuribenzenesulfonic acid. Brain Res. 1992;591:341–344. doi: 10.1016/0006-8993(92)91717-s. [DOI] [PubMed] [Google Scholar]
  112. Wyatt CN, Peers C. Ca2+-activated K+ channels in isolated type I cells of the neonatal rat carotid body. J Physiol. 1995;483:559–565. doi: 10.1113/jphysiol.1995.sp020606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Wyatt CN, Peers C, Kumar P, Hardie DG, Evans AM. The effect of AMP-kinase activators on oxygen-sensing type 1 cells of the rat carotid body. pA2 online. 2006 http://www.pa2online.org. [Google Scholar]
  114. Youngson C, Nurse C, Yeger H, Katz E. Oxygen sensing in airway chemoreceptors. Nature. 1993;356:153–155. doi: 10.1038/365153a0. [DOI] [PubMed] [Google Scholar]
  115. Zhao Y, Packer CS, Rhoades RA. The vein utilizes different sources of energy than the artery during pulmonary hypoxic vasoconstriction. Exp Lung Res. 1996;22:51–63. doi: 10.3109/01902149609074017. [DOI] [PubMed] [Google Scholar]
  116. Zhou G, Myers R, Li Y, Chen Y, Shen X, Fenyk-Melody J, Wu M, Ventre J, Doebber T, Fujii N, Musi N, Hirshman MF, Goodyear LJ, Moller DE. Role of AMP-activated protein kinase in mechanism of metformin action. J Clin Invest. 2001;108:1167–1174. doi: 10.1172/JCI13505. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from The Journal of Physiology are provided here courtesy of The Physiological Society

RESOURCES