Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Adaptation to lowering oxygen levels (hypoxia) requires coordinated downregulation of metabolic demand and supply to prevent a mismatch in ATP utilization and production that might culminate in a bioenergetic collapse. Hypoxia diminishes ATP utilization by downregulating protein translation and the activity of the Na-K-ATPase. Hypoxia diminishes ATP production in part by lowering the activity of the electron transport chain through activation of the transcription factor hypoxia-inducible factor-1. The decrease in electron transport limits the overproduction of reactove oxygen species during hypoxia and slows the rate of oxygen depletion to prevent anoxia. In this review, we discuss these mechanisms that diminish metabolic supply and demand for adaptation to hypoxia.

Free full text 


Logo of ajpcellLink to Publisher's site
Am J Physiol Cell Physiol. 2011 Mar; 300(3): C385–C393.
Published online 2010 Dec 1. https://doi.org/10.1152/ajpcell.00485.2010
PMCID: PMC3063979
PMID: 21123733

Hypoxia. 2. Hypoxia regulates cellular metabolism

Abstract

Adaptation to lowering oxygen levels (hypoxia) requires coordinated downregulation of metabolic demand and supply to prevent a mismatch in ATP utilization and production that might culminate in a bioenergetic collapse. Hypoxia diminishes ATP utilization by downregulating protein translation and the activity of the Na-K-ATPase. Hypoxia diminishes ATP production in part by lowering the activity of the electron transport chain through activation of the transcription factor hypoxia-inducible factor-1. The decrease in electron transport limits the overproduction of reactove oxygen species during hypoxia and slows the rate of oxygen depletion to prevent anoxia. In this review, we discuss these mechanisms that diminish metabolic supply and demand for adaptation to hypoxia.

Keywords: sodium-potassium-ATPase, electron transport chain, oxygen levels

metazoans have developed mechanisms for adaptation to lower oxygen levels (hypoxia) (97). One of these mechanisms is metabolic suppression as measured by a decrease in the mitochondrial oxygen consumption of cells (i.e., the respiratory rate) occurring at oxygen levels (1–3% O2) well above the threshold where oxygen becomes limiting (<0.3% O2) to cytochrome c oxidase (COX) (complex IV), a phenomenon referred to as oxygen conformance (53). Complex IV is the main enzyme in the electron transport chain (ETC) that utilizes oxygen and couples it to the generation of ATP (oxidative phosphorylation). Based on the ability of deep sea diving mammals to survive long periods of hypoxia, Peter Hochachaka (52) elegantly surmised that animals tolerant of hypoxia have the ability to undergo oxygen conformance as an adaptive mechanism. Furthermore, studies of intact myocardium indicated that regional decreases in oxygen delivery can elicit local decreases in oxygen demand as reflected by a diminished contractile function (2, 3, 75, 88). However, most studies in isolated cells failed to observe oxygen conformance under hypoxic conditions (6, 8, 109), even though metabolic suppression was observed during anoxia (0% O2) (5, 7). Because this metabolic suppression occurred at oxygen levels where COX is inhibited, the oxygen conformance phenomenon remained theoretical. In the 1990s, multiple studies observed that isolated cells are able to diminish the respiratory rate at oxygen levels in the range of 1–3% O2 by decreasing the cellular ATP-utilizing processes (metabolic demand) (22, 23, 31, 96). Previous studies failed to observe a decrease in the respiratory rate during hypoxia because isolated cells were exposed to hypoxia for seconds (8, 109). However, when cells are exposed to chronic hypoxia (minutes to hours), they displayed a reversible suppression of oxygen consumption without any detectable cell injury (22, 23, 31, 96). In recent years, it has been observed that hypoxic suppression of the respiratory rate also involves the activation of the transcription factor hypoxia-inducible factor (HIF-1), which regulates ATP generation (metabolic supply) (98). In this review we focus on the mechanisms underlying the coordinated regulation of metabolic demand and supply during hypoxia resulting in metabolic adaptation to maintain cellular homeostasis.

What Controls the Respiratory Rate of Cells?

To understand how hypoxia causes a decrease in the respiratory rate, we briefly review the factors that are responsible for the control of respiration. The respiratory rate is the rate of oxygen consumption by mitochondria in living cells. Although mitochondria consume the majority of cellular oxygen, there are other cellular processes that also consume oxygen. It is important to separate nonmitochondrial oxygen consumption from mitochondrial oxygen consumption in living cells by measuring oxygen consumption in the presence of mitochondrial inhibitors, such as the complex I and III inhibitors rotenone and antimycin, respectively. Mitochondrial respiration proceeds when reducing equivalents (NADH and FADH2) are generated by the tricarboxylic acid (TCA) cycle (93). The electrons generated from NADH and FADH2 undergo oxidation to NAD+ and FAD+ by complex I and II of the ETC located in the inner mitochondrial membrane. Subsequently, these electrons are sequentially transferred to complex III, cytochrome c, and complex IV, which transfers the electrons to molecular oxygen. The movement of electrons through the ETC is coupled to proton translocation from the mitochondrial matrix across the inner membrane to the intermitochondrial membrane space, creating an electrochemical gradient of protons consisting of a pH gradient and a membrane potential. These protons can either run down their gradient through the F1Fo-ATP synthase (complex V) or the protons can leak back across the inner membrane to the mitochondrial matrix (14). Complex V couples the transport of protons to the generation of ATP from ADP and phosphate (Pi). ATP generated in the matrix is exported to the cytosol in exchange for ADP by the adenine nucleotide translocase (ANT) located in the inner membrane. ATP in the cytosol is utilized by variety of processes such as the Na-K-ATPase, which regenerates ADP pools. ADP is then transported back to the mitochondria, so the cycle can continue. Thus mitochondrial oxygen consumption is a combination of coupled respiration and uncoupled respiration. Coupled respiration is the rate of oxygen consumption by complex IV coupled to the generation of ATP synthesis by complex V. By contrast, uncoupled respiration is the rate of oxygen consumption by complex IV that is not coupled to the generation of ATP due to the proton leak. Most cells display high levels of coupled respiration. The notable exception are brown fat cells that exhibit uncoupled respiration due to an abundance of uncoupling proteins, which increase proton leak by allowing protons to flow back into mitochondria without driving the generation of ATP (40).

Initial work in the 1950s by Chance and Williams (29) proposed that the respiratory rate in cells is controlled by cellular ATP utilization. In this model, the increase in cellular ATP utilization decreases cytosolic ATP levels and increases cytosolic ADP and Pi levels. The rise in cytosolic ADP levels leads to a rise in mitochondrial ADP via the increased activity of the adenine nucleotide carrier. The increased mitochondrial ADP concentration stimulates the ATP synthase to augment the rate of ATP synthesis, which results in a decrease in the mitochondrial membrane potential, thus stimulating the respiratory chain to consume oxygen. However, in the ensuing decades it has become clear that other factors also control the respiratory rate, such as the availability of reducing equivalents provided by the TCA cycle, electron flux through the ETC, the availability of ADP provided by cellular ATPases, the adenine nucleotide translocase, and the magnitude of the proton leak (64). How to quantitatively measure the relative control exerted by these processes on the respiratory rate remained elusive until the pioneering work of Kacser and Burns (66) in the early 1970s describing metabolic control analysis. Their work describing metabolic control analysis helped them determine control coefficients of a particular protein over metabolic flux through a pathway. The control coefficient is the percent change in the respiratory rate divided by the percent change in the protein or complex causing the change in the respiratory rate. For example, if a 10% change in the ANT results in a 10% change in the respiratory rate, then the control coefficient of the proton leak would be 1. However, if the 10% change resulted in 1% change in the respiratory rate, then the control coefficient would be 0.1. In the late 1980s, Brand and colleagues began to utilize metabolic control analysis on isolated rat hepatocytes exposed to ambient air and determined that 15–30% of respiration is controlled by the NADH supply (these include pyruvate supply to the mitochondria, the TCA cycle, and any other NADH-supplying reaction); 20% is controlled by the proton leak; and 0–15% is controlled by the ETC (16, 17). The remaining 50% is controlled by ATP synthesis, transport, and utilization of which in most cells the dominant factor is the rate of ATP utilization by cellular processes. Chandel et al. (31) performed metabolic control analysis in isolated rat hepatocytes and determined that there was no major difference between normoxia and hypoxia with respect to factors that controlled the respiratory rate-specifically NADH supply, ETC activity, and ATP utilization.

Hypoxia Diminishes NADH Supply to the ETC Through HIF-1

Under normal oxygen conditions, pyruvate derived from glycolysis is transported into the mitochondria and converted into acetyl-CoA by the pyruvate dehydrogenase (PDH) complex (112). Acetyl-CoA combines with oxaloacetate to form citrate as the initial step in the TCA cycle. The reducing equivalents NADH and FADH2 generated by the TCA cycle drive ETC to generate ATP for energy and ROS for signaling, while the TCA cycle intermediates are utilized for biosynthetic processes such as lipid synthesis (107). During acute exposure to hypoxia (seconds to minutes), cells do not display changes in carbon flux into the TCA cycle. However, as cells are exposed to prolonged hypoxia (hours), the activation of the transcription factor HIF-1 diminishes the carbon flux into the TCA cycle (71, 82, 89). HIF-1 is a heterodimer of two basic helix loop-helix/PER-SIM-ARNT(PAS) proteins: HIF-1α and the aryl hydrocarbon nuclear trans-locator (ARNT or HIF-1β) (104). Both HIF-1α and HIF-1β protein subunits are expressed ubiquitously, but the stability of each protein is differentially regulated by oxygen levels (104). HIF-1α protein is rapidly degraded under normal oxygen conditions, whereas HIF-1β protein levels are constitutively stable. During normoxia, HIF-1α protein undergoes hydroxylation at two proline residues 402 and 564 within the oxygen-dependent degradation domain of HIF-1α (61, 63, 83). This hydroxylation reaction is catalyzed by prolyl hydroxylases (PHDs) that require Fe (II), oxygen, and 2-oxoglutarate (21, 41). Hydroxylated proline serves as a binding site for the von Hippel-Lindau protein (pVHL), the substrate recognition component of the VBC-CUL-2 E3 ubiquitin ligase complex (55). Once bound, pVHL tags HIF-1α with ubiquitin thereby targeting it for proteasomal degradation (84). Hypoxia prevents hydroxylation of HIF-1α protein resulting in the stabilization of the protein. Hypoxia diminishes hydroxylation collectively by 1) limiting the oxygen as a substrate (67) and 2) through an increase in the release of superoxide from mitochondrial complex III (10) (Fig. 1). The mechanisms linking mitochondrially generated superoxide with the stabilization of the HIF are an area of active investigation. According to one hypothesis, complex III-generated superoxide is converted to hydrogen peroxide in the cytosol where it oxidizes the PHDs cofactor Fe (II) to Fe (III) making the PHDs catalytically inactive (45).

An external file that holds a picture, illustration, etc.
Object name is zh00021165190001.jpg

Hypoxia activates hypoxia inducible factor-1 (HIF-1). HIF1α subunit is hydroxylated by pyruvate dehydrogenase 2 (PHD2) at distinct proline residues thereby targeting the protein for von Hippel-Lindau protein (pVHL)-mediated proteasomal degradation. Hypoxia concomitantly diminishes PHD2 activity and induces the production of mitochondrial reactive oxygen species (ROS) at complex III resulting in an inhibition of hydroxylation of HIF1α subunit. Once HIF1α subunit is stabilized, it binds with HIF-β and p300 coactivators to hypoxic response elements (HREs) in the promoters and enhancers of target genes that modulate metabolism.

The two major targets of HIF-1 that decrease pyruvate conversion to acetyl-coA are lactate dehydrogenase A (LDH-A) and pyruvate dehydrogenase kinase 1 (PDK1) (43, 71, 89). PDK1 phosphorylates and inactivates the catalytic subunit of PDH (54). An increase in PDK1 protein levels decreases PDH activity thereby preventing the conversion of pyruvate to acetyl-coA while also driving pyruvate conversion to lactate. LDH-A converts pyruvate to lactate by utilizing the NADH generated from glycolysis (103). This is an efficient manner to oxidize cytosolic NADH to NAD+, which allow for an increase in the glycolytic rate because the regeneration of NAD+ is a rate-limiting step for glycolysis. Normally, reducing equivalents from cytosolic NADH are shuttled into the mitochondria via the malate aspartate shuttle where NADH is regenerated to NAD+ and subsequently shuttled back to the cytosol. The ectopic expression of LDH-A alone is sufficient to increase lactate production (102). The coordinated upregulation of PDK1 and LDH-A diverts pyruvate from fueling the mitochondria to generation of lactate (Fig. 2). The decreased acetyl-coA levels in the mitochondria diminish TCA cycle activity, resulting in reduced generation of mitochondrial NADH and FADH2 levels and reduced electron flux through the ETC. Based on this model, we hypothesize that cells that are efficient in conducting fatty acid oxidation to generate mitochondrial acetyl-coA would be refractory to PDK1/LDH-A dependent decrease in TCA cycle activity. We predict that only cells that utilize pyruvate as the major carbon source to fuel the TCA cycle would be sensitive to inhibition of the TCA cycle by the PDK1/LDH-A axis.

An external file that holds a picture, illustration, etc.
Object name is zh00021165190002.jpg

HIF-1 shifts metabolism from oxidative to glycolysis. HIF-1 promotes the activation of glycolysis by upregulating numerous glycolytic genes including lactate dehydrogenase A (LDH-A), which converts pyruvate to lactate. Pyruvate conversion into acetyl-coA is dependent on PDH. HIF-1 also induces PDK1, a negative regulator of PDH. Thus HIF-1 diminishes pyruvate entry into the tricarboxylic acid (TCA) cycle and oxidative metabolism.

Hypoxia Diminishes Electron Transport

Multiple studies throughout the 1970s and 1980s examined the oxygen dependence of the ETC (108, 109). These studies observed that exposure of cells to acute hypoxia (minutes to seconds) did not attenuate the flux of electrons through the ETC nor increase NADH levels in mitochondria. However, in the mid-1990s we reported that isolated mitochondria decreased coupled respiration and that isolated COX decreased its maximal velocity (Vmax) when exposed to chronic hypoxia (2 h) (30, 32). Thus there is an intrinsic oxygen dependence of COX during prolonged hypoxia. Another important regulator of COX activity is nitric oxide (NO) (36). Low concentrations (nM range) of NO reversibly inhibit isolated COX by competing with oxygen (15, 35, 80). Under aerobic conditions, oxygen levels are high enough to prevent NO from inhibiting COX activity (36). However, as oxygen levels fall, the low levels of NO are sufficient to inhibit COX activity. Low levels of NO under normoxia do not injure cells. However, the same low levels of NO are sufficient to inhibit respiration and initiate cell death under hypoxia (1.5% O2) (76). In the absence of NO, hypoxia alone does not have any deleterious effects on cells. It is likely that COX activity is compromised in inflammatory conditions where NO levels are high with concomitant tissue hypoxia. Furthermore, the NO-generating enzyme inducible NO synthase (iNOS) is a target of HIF-1 (65, 85). We propose that hypoxia diminishes COX activity by decreasing the Vmax of COX activity and by increasing NO levels to inhibit COX activity. Although this mechanism diminishes COX activity during hypoxia, the activity cannot be diminished to the point where respiration fails to meet the basal metabolic demands of cells. Therefore, cells ensure optimal COX activity during hypoxia by activating HIF-1 to induce subunit switch from COX4–1 subunit to COX4–2 (44). COX has 13 subunits, of which the three catalytic subunits COX I-III are encoded by mitochondrial DNA. The remaining regulatory 10 subunits including COX4 subunits are encoded by nuclear DNA. HIF-1 induces both the expression of the COX4–2 subunit and the mitochondrial protease LON, which targets COX4–1 subunit degradation to complete the switching of the COX4 subunits during hypoxia. Recently, another mechanism to downregulate the ETC is the finding that micro-RNA 210 (mir-210) blocks the expression of the iron-sulfur cluster assembly proteins ISCU1/2, which are required for the functions of complex I, COX subunit 10, aconitase, and subunit D of succinate dehydrogenase (28, 33, 42, 91). Using a miRNA microarray, Kulshreshta et al. (74) first discovered that miR-210 is regulated by hypoxia, and recently it was proposed to be the major micro-RNA upregulated during hypoxia. HIF-1, but not HIF-2, is responsible for the induction of mir-210 during hypoxia (57). The ectopic expression of mir-210 is sufficient to decrease mitochondrial respiration and upregulate glycolysis (33). Thus there are multiple mechanisms by which HIF-1 can coordinately diminish electron flux through the ETC (Fig. 3).

An external file that holds a picture, illustration, etc.
Object name is zh00021165190003.jpg

Hypoxia diminishes electron flux through the electron transport chain. Hypoxia diminishes respiratory activity by activating HIF-1, which increases micro-RNA 210 (miR-210), inducible nitric oxide synthase (iNOS), and switching of cytochrome c oxidase (COX)4–1 subunit to COX4–2. Hypoxia can also directly decrease complex IV (COX) activity.

Hypoxia Diminishes Cellular ATP Utilization

A major ATP consumer is the Na-K-ATPase, which can account for 20–70% of the oxygen expenditure of mammalian cells (86). Na-K-ATPase is a plasma membrane protein that transports Na+ and K+ across the plasma membrane to maintain ionic gradients (69). The Na-K-ATPase is a heterodimer composed of α- and β-subunits. The α-subunit is a transmembrane protein that cleaves high-energy phosphate bonds and exchanges intracellular Na+ for extracellular K+ by coupling the exchange to the hydrolysis of ATP. The smaller β-subunit is a glycosylated transmembrane protein that controls the heterodimer assembly and insertion into the plasma membrane. Multiple investigators have reported that hypoxia reversibly suppresses Na-K-ATPase activity (27, 38, 81, 90, 110). Initially, the Na-K-ATPase catalytic activity was proposed to be regulated through changes in substrate affinity. However, recent reports have demonstrated that the Na-K-ATPase activity is regulated by phosphorylation, which results in either endocytosis or exocytosis of this molecule from the plasma membrane to internal compartments (34, 92). Indeed, exposure to hypoxia for as little as 15 min decreases Na-K-ATPase activity due to endocytosis of the α-subunit from the plasma membrane (37). AMP-activated protein kinase (AMPK) directly phosphorylates PKC-ζ at Thr410 to promote Na-K-ATPase endocytosis during hypoxia (46) (Fig. 4). Whereas hypoxia does not affect the total Na-K-ATPase protein abundance in cell lysates during this short exposure (<1 h), chronic exposure causes hypoxia-induced endocytosis of the Na-K-ATPase triggering pVHL-mediated degradation of plasma membrane Na-K-ATPase in a HIF independent manner (113).

An external file that holds a picture, illustration, etc.
Object name is zh00021165190004.jpg

Hypoxia diminishes cellular ATP demand. Protein translation and Na-K-ATPase are two major ATP-utilizing processes in the cell. Hypoxia through the generation of mitochondrial ROS activates AMP-activated protein kinase (AMPK) resulting in a decrease in Na-K-ATPase activity and inhibition of mammalian target of rapamycin (mTORC1)-dependent translation. Hypoxia also activates pancreatic eIF2α kinase (PERK) to dampen protein translation.

The other major consumer of ATP is mRNA translation to protein. Multiple studies have demonstrated that hypoxia inhibits mRNA translation (4, 1113, 58, 59, 72, 73, 77). The initiation of mRNA translation is regulated by the active eukaryotic initiation factors eIF4F and eIF2 (51). The suppression of mammalian target of rapamycin (mTOR) complex 1 (mTORC1) and pancreatic eIF2α kinase (PERK) are the key regulators of diminished global translation observed during hypoxia (78). mTORC1 consists of the catalytic subunit mTOR and several regulatory proteins: raptor, PRS40, DEPTOR, and mLST8/GβL (99). mTORC1 controls translation by phosphorylating the p70 ribosomal S6 kinase (S6K1) and 4E-BP1, the eukaryotic initiation factor 4E (eIF4E) binding protein 1. mTORC1 is activated by the small Ras-like GTPase Rheb, which is inhibited by TSC2. During growth factor stimulation, AKT phosphorylates and inhibits TSC2 thereby allowing Rheb to activate mTORC1. Hypoxia causes rapid (within 15 min) and reversible hypophosphorylation of mTORC1 effectors 4E-BP1 and S6K (4, 58). The rapid inhibition of mTOR is HIF independent and occurs through the activation of AMPK (4, 77). AMPK phosphorylates TSC2 and raptor resulting in suppression of mTORC1 (100). The loss of TSC2 effectively suppresses mTOR inhibition during hypoxia (20). The sustained inhibition of mTORC1 over hours involves the HIF-dependent transcription of REDD1, which releases TSC2 from its growth factor-stimulated association with 14–3-3 proteins to allow TSC2 inhibition of mTORC1 (19). AMPK can also phosphorylate eEF2 kinase (eEF2K) resulting in eEF2 phosphorylation and inhibition of translation elongation (18, 56). Another negative regulator of mTORC1 during hypoxia is the promyelocytic leukemia protein, which promotes sequestration of mTOR to the nucleus during hypoxia (11). Finally, mRNA translation during hypoxia is decreased by activation of PERK (12, 13, 72, 73). PERK activation results in eIF2α phosphorylation, which inhibits mRNA translation initiation (95). PERK activation during hypoxia requires mitochondrial ROS (79) (Fig. 4). Collectively, these studies highlight that multiple mechanisms contribute to the decrease in global protein synthesis during hypoxia.

It is becoming increasingly clear that hypoxic activation of AMPK diminishes processes that consume ATP such as Na-K-ATPase and global protein synthesis allowing metabolic adaptation to decreasing oxygen levels. AMPK is a heterotrimeric serine/threonine consisting of a catalytic α subunit and two regulatory β- and γ-subunits (68). AMPK is ubiquitously expressed and is activated upon nutrient-limiting conditions. AMPK has evolved to serve as a metabolic checkpoint in cells by halting cell growth and suppressing ATP-consuming processes (48). AMPK is activated either by an increase in the AMP-to-ATP ratio through the protein kinase LKB1 or an increase in calcium through the Ca2+/calmodulin-dependent kinase kinase (CaMKK) (49, 50, 101, 111). Multiple studies have demonstrated that acute hypoxia (minutes to several hours) does not alter AMP-to-ATP ratio but does robustly activate AMPK (39, 77). This is not surprising since oxygen becomes limiting to the generation of ATP as cells approach anoxic conditions. Since COX is not limited under hypoxic conditions, there is no reason for cells to experience a decrease in the ability to generate ATP during acute hypoxia (32). As it is well established that cells under hypoxia elicit an increase in calcium (105), the more likely scenario is that CaMKK is the upstream kinase required for hypoxic activation of AMPK. However, the experimental evidence for CaMKK-dependent hypoxic activation of AMPK is still lacking. It is also possible that LKB1 activates AMPK independent of changes in the AMP-to-ATP ratio during hypoxia. Interestingly, the increase in calcium during hypoxia has been suggested to be under the control of mitochondrial complex III-generated ROS (106). We recently demonstrated that rapid activation of AMPK during hypoxia is dependent on mitochondrial complex III ROS (39). This is consistent with the observations that the suppression of ATP-consuming processes is dependent on mitochondrial ROS generation. Therefore, metabolic demand during hypoxia is likely initiated by mitochondrial ROS activation of AMPK (Fig. 4).

Physiological Consequences of Metabolic Suppression

An advantage of cells reducing their respiratory rate as oxygen delivery diminishes would be to delay the onset of local tissue anoxia. Furthermore, an attenuated respiratory rate might render cells less susceptible to anoxia-induced cell injury. Enhanced resistance to anoxic injury might develop if the adapted cells could selectively inhibit facultative metabolic activities related to organ function, thereby preserving limited energy production for obligatory functions necessary for cell survival (52). Metabolic suppression is observed in intact hearts where regional decreases in myocardial oxygen delivery results in decreased contractile activity and oxygen consumption, a phenomenon termed hibernating myocardium (2, 3, 75, 88). In the early 1990s, Bristow and colleagues (2, 3) observed diminished contraction in swine myocardium due to a reduction in epicardial flow. Downey and Lee (75) also found evidence of metabolic suppression in isoprenaline-stimulated canine myocardium during partial ischemia. Importantly, no detectable cellular injury was observed, and contractile function was recovered upon restoration of blood flow to the ischemic regions. Thus hibernating myocardium represents a state where a reduction in ATP demand in response to a decrease in regional oxygen supply serves to protect the myocardium from developing ischemic injury when the blood flow is severely reduced.

The other major advantage of diminishing respiratory rate during hypoxia is to limit the production of ROS (98). Within minutes, hypoxia increases ROS production from complex III to stabilize the HIF-1α protein (10). ROS at low levels initiate cellular signaling events but at high levels can initiate damage and cell death (47). Interestingly, cells maintained for several days under hypoxia (chronic hypoxia) decrease the levels of ROS back to normoxic levels (71). However, HIF-1 knockout cells continue to display increased ROS during hypoxia resulting in cell death, and the addition of ROS scavengers decrease hypoxia-induced cell death in HIF-1 knockout cells indicating that HIF-1 dampening of ROS levels is an adaptive mechanism (71). Furthermore, expression of PDK1 in HIF-1-deficient cells reduces ROS levels and apoptosis during hypoxia (71). COX4–2 subunit switching to COX4–1 subunit during hypoxia also affects ROS production, as hypoxic cells in which COX4–2 is diminished but still express COX4–1 display increased ROS levels and apoptosis (44). The expression of mir210 also decreases ROS generation during hypoxia by decreasing ETC activity (28, 33, 42). Collectively, these in vitro data indicate that reduced flux of carbons through the TCA cycle and subsequent electron flux through the ETC during hypoxia is an adaptive response to diminish overproduction of ROS (Fig. 5).

An external file that holds a picture, illustration, etc.
Object name is zh00021165190005.jpg

Hypoxia depresses the respiratory rate for metabolic adaptation. The downregulation of ATP demand and supply diminishes the respiratory rate, which prevents the overproduction of ROS and depletion of oxygen under hypoxic conditions.

Presently, the full in vivo significance of HIF-1-mediated decrease in ROS production is not fully understood. To date, the best evidence suggesting that HIF-1-mediated depression of the ETC and ROS generation is relevant in vivo comes from studies on ischemia-reperfusion injury. Tissues exposed to ischemia followed by restoration of blood supply undergo oxidative stress-induced damage called reperfusion injury. Multiple studies have demonstrated that reducing flux through the ETC during ischemia and reperfusion phases prevents reperfusion injury (24). If reducing ETC flux prevents injury, then one would predict that activation of HIF-1 repression of the TCA cycle and the ETC would prevent reperfusion injury. To test this possibility Semenza and colleagues (62) have utilized heterozygous HIF-1α (HIF1α+/−) mice because homozygous HIF-1α mice are embryonic lethal. They observed that exposure of wild-type mice to intermittent hypoxia resulted in protection of isolated hearts against ischemia-reperfusion injury 24 h later (26) and that this cardiac protection was lost in HIF1α+/− mice (25). Subsequently, multiple studies have demonstrated that activation of HIF-1 pharmacologically using PHD inhibitors or genetically through small interfering RNA (siRNA) against PHDs or via constitutive expression of HIF-1α also prevents ischemia-reperfusion injury in the heart (9, 70, 87). PHD2 hypomorph mice are also protected against ischemia-reperfusion injury in the heart (60). Recent studies using PHD1 null mice further corroborate these studies. Aragones et al. (1) observed that ablation of PHD1 results in increased in expression of PDK1 and PDK4 leading to repression of mitochondrial metabolism. PHD1 null mice were found to be remarkably protected against ischemia-reperfusion injury in the liver. Short-term inhibition of PHD1 through siRNA also prevents liver ischemia-reperfusion injury (94). The loss of PHD1 production correlated with a decrease in oxidative stress due to depression of mitochondrial metabolism. However, it is not yet clear whether these HIF-1-induced cytoprotective effects observed against ischemia-reperfusion injury are primarily due to diminished mitochondrial metabolism and ROS generation as HIF-1 can also activate a variety of other protective mechanisms such as the production of the hormone erythropoietin (EPO), which is sufficient to prevent ischemia-reperfusion injury (25). It will be interesting to test whether HIF-1-mediated protection against ischemia-reperfusion injury acts through the inhibition of mitochondrial metabolism and ROS generation through induction of mir-210 or PDK1.

In conclusion, there is an ongoing lively debate in the bioenergetic field over what controls the rate of cellular respiration during hypoxia. Historically, investigators attributed the majority of metabolic control to the activity of cellular ATPases, which inhibit respiration by decreasing the availability of ADP. Recent work from several groups of investigators has challenged this model, suggesting that reducing the supply of carbon to the TCA cycle and reducing electron flux through the ETC can also inhibit respiration. This debate is roughly analogous to the ongoing debate over what controls the United States economy; one camp argues for a strategy that increases demand through public sector spending while the other argues for a strategy that increases the money supply through lowered interest and tax rates. Currently, available data suggests that the cell employs strategies regulating both supply and demand simultaneously to reduce metabolism during hypoxia to maintain homeostasis. The downregulation of ATP utilization by bioenergetic processes, including protein synthesis and Na-K-ATPase activity, is coordinated with a reduction in carbon flux through the TCA cycle and electron flux through the ETC by activating HIF-1-mediated induction of PDK1 and mir-210 as well as subunit switching of COX subunit 4. This coordinated downregulation of demand and supply prevents a mismatch in ATP production and utilization that might culminate in bioenergetic collapse while limiting the production of ROS and slowing the rate of oxygen depletion under ischemic conditions (Fig. 5).

GRANTS

This work was supported by a National Institutes of Health Grant R01CA123067-04 to N. S. Chandel, and W. W. Wheaton was supported by a predoctoral training grant 5T32CA009560-24.

DISCLOSURES

No conflicts of interest, financial or otherwise, are declared by the author(s).

ACKNOWLEDGMENTS

Because of the brevity of this review, we apologize to those whose important work we could not discuss.

REFERENCES

1. Aragones J, Schneider M, Van Geyte K, Fraisl P, Dresselaers T, Mazzone M, Dirkx R, Zacchigna S, Lemieux H, Jeoung NH, Lambrechts D, Bishop T, Lafuste P, Diez-Juan A, Harten SK, Van Noten P, De Bock K, Willam C, Tjwa M, Grosfeld A, Navet R, Moons L, Vandendriessche T, Deroose C, Wijeyekoon B, Nuyts J, Jordan B, Silasi-Mansat R, Lupu F, Dewerchin M, Pugh C, Salmon P, Mortelmans L, Gallez B, Gorus F, Buyse J, Sluse F, Harris RA, Gnaiger E, Hespel P, Van Hecke P, Schuit F, Van Veldhoven P, Ratcliffe P, Baes M, Maxwell P, Carmeliet P. Deficiency or inhibition of oxygen sensor Phd1 induces hypoxia tolerance by reprogramming basal metabolism. Nat Genet 40: 170–180, 2008 [Abstract] [Google Scholar]
2. Arai AE, Grauer SE, Anselone CG, Pantely GA, Bristow JD. Metabolic adaptation to a gradual reduction in myocardial blood flow. Circulation 92: 244–252, 1995 [Abstract] [Google Scholar]
3. Arai AE, Pantely GA, Anselone CG, Bristow J, Bristow JD. Active downregulation of myocardial energy requirements during prolonged moderate ischemia in swine. Circ Res 69: 1458–1469, 1991 [Abstract] [Google Scholar]
4. Arsham AM, Howell JJ, Simon MC. A novel hypoxia-inducible factor-independent hypoxic response regulating mammalian target of rapamycin and its targets. J Biol Chem 278: 29655–29660, 2003 [Abstract] [Google Scholar]
5. Aw TY, Andersson BS, Jones DP. Suppression of mitochondrial respiratory function after short-term anoxia. Am J Physiol Cell Physiol 252: C362–C368, 1987 [Abstract] [Google Scholar]
6. Aw TY, Jones DP. ATP concentration gradients in cytosol of liver cells during hypoxia. Am J Physiol Cell Physiol 249: C385–C392, 1985 [Abstract] [Google Scholar]
7. Aw TY, Jones DP. Secondary bioenergetic hypoxia. Inhibition of sulfation and glucuronidation reactions in isolated hepatocytes at low O2 concentration. J Biol Chem 257: 8997–9004, 1982 [Abstract] [Google Scholar]
8. Aw TY, Wilson E, Hagen TM, Jones DP. Determinants of mitochondrial O2 dependence in kidney. Am J Physiol Renal Fluid Electrolyte Physiol 253: F440–F447, 1987. [Abstract] [Google Scholar]
9. Bao W, Qin P, Needle S, Erickson-Miller CL, Duffy KJ, Ariazi JL, Zhao S, Olzinski AR, Behm DJ, Pipes GC, Jucker BM, Hu E, Lepore JJ, Willette RN. Chronic inhibition of hypoxia-inducible factor prolyl 4-hydroxylase improves ventricular performance, remodeling, and vascularity after myocardial infarction in the rat. J Cardiovasc Pharmacol 56: 147–155 [Abstract] [Google Scholar]
10. Bell EL, Klimova TA, Eisenbart J, Moraes CT, Murphy MP, Budinger GR, Chandel NS. The Qo site of the mitochondrial complex III is required for the transduction of hypoxic signaling via reactive oxygen species production. J Cell Biol 177: 1029–1036, 2007 [Europe PMC free article] [Abstract] [Google Scholar]
11. Bernardi R, Guernah I, Jin D, Grisendi S, Alimonti A, Teruya-Feldstein J, Cordon-Cardo C, Simon MC, Rafii S, Pandolfi PP. PML inhibits HIF-1alpha translation and neoangiogenesis through repression of mTOR. Nature 442: 779–785, 2006 [Abstract] [Google Scholar]
12. Bi M, Naczki C, Koritzinsky M, Fels D, Blais J, Hu N, Harding H, Novoa I, Varia M, Raleigh J, Scheuner D, Kaufman RJ, Bell J, Ron D, Wouters BG, Koumenis C. ER stress-regulated translation increases tolerance to extreme hypoxia and promotes tumor growth. EMBO J 24: 3470–3481, 2005 [Europe PMC free article] [Abstract] [Google Scholar]
13. Blais JD, Filipenko V, Bi M, Harding HP, Ron D, Koumenis C, Wouters BG, Bell JC. Activating transcription factor 4 is translationally regulated by hypoxic stress. Mol Cell Biol 24: 7469–7482, 2004 [Europe PMC free article] [Abstract] [Google Scholar]
14. Boyer PD. The ATP synthase–a splendid molecular machine. Annu Rev Biochem 66: 717–749, 1997 [Abstract] [Google Scholar]
15. Brown GC, Cooper CE. Nanomolar concentrations of nitric oxide reversibly inhibit synaptosomal respiration by competing with oxygen at cytochrome oxidase. FEBS Lett 356: 295–298, 1994 [Abstract] [Google Scholar]
16. Brown GC, Hafner RP, Brand MD. A ‘top-down’ approach to the determination of control coefficients in metabolic control theory. Eur J Biochem 188: 321–325, 1990 [Abstract] [Google Scholar]
17. Brown GC, Lakin-Thomas PL, Brand MD. Control of respiration and oxidative phosphorylation in isolated rat liver cells. Eur J Biochem 192: 355–362, 1990 [Abstract] [Google Scholar]
18. Browne GJ, Finn SG, Proud CG. Stimulation of the AMP-activated protein kinase leads to activation of eukaryotic elongation factor 2 kinase and to its phosphorylation at a novel site, serine 398. J Biol Chem 279: 12220–12231, 2004 [Abstract] [Google Scholar]
19. Brugarolas J, Lei K, Hurley RL, Manning BD, Reiling JH, Hafen E, Witters LA, Ellisen LW, Kaelin WG., Jr Regulation of mTOR function in response to hypoxia by REDD1 and the TSC1/TSC2 tumor suppressor complex. Genes Dev 18: 2893–2904, 2004 [Europe PMC free article] [Abstract] [Google Scholar]
20. Brugarolas JB, Vazquez F, Reddy A, Sellers WR, Kaelin WG., Jr TSC2 regulates VEGF through mTOR-dependent and -independent pathways. Cancer Cell 4: 147–158, 2003 [Abstract] [Google Scholar]
21. Bruick RK, McKnight SL. A conserved family of prolyl-4-hydroxylases that modify HIF. Science 294: 1337–1340, 2001 [Abstract] [Google Scholar]
22. Budinger GR, Chandel N, Shao ZH, Li CQ, Melmed A, Becker LB, Schumacker PT. Cellular energy utilization and supply during hypoxia in embryonic cardiac myocytes. Am J Physiol Lung Cell Mol Physiol 270: L44–L53, 1996 [Abstract] [Google Scholar]
23. Budinger GR, Duranteau J, Chandel NS, Schumacker PT. Hibernation during hypoxia in cardiomyocytes. Role of mitochondria as the O2 sensor. J Biol Chem 273: 3320–3326, 1998 [Abstract] [Google Scholar]
24. Burwell LS, Brookes PS. Mitochondria as a target for the cardioprotective effects of nitric oxide in ischemia-reperfusion injury. Antioxid Redox Signal 10: 579–599, 2008 [Abstract] [Google Scholar]
25. Cai Z, Manalo DJ, Wei G, Rodriguez ER, Fox-Talbot K, Lu H, Zweier JL, Semenza GL. Hearts from rodents exposed to intermittent hypoxia or erythropoietin are protected against ischemia-reperfusion injury. Circulation 108: 79–85, 2003 [Abstract] [Google Scholar]
26. Cai Z, Zhong H, Bosch-Marce M, Fox-Talbot K, Wang L, Wei C, Trush MA, Semenza GL. Complete loss of ischaemic preconditioning-induced cardioprotection in mice with partial deficiency of HIF-1 alpha. Cardiovasc Res 77: 463–470, 2008 [Abstract] [Google Scholar]
27. Carpenter TC, Schomberg S, Nichols C, Stenmark KR, Weil JV. Hypoxia reversibly inhibits epithelial sodium transport but does not inhibit lung ENaC or Na-K-ATPase expression. Am J Physiol Lung Cell Mol Physiol 284: L77–L83, 2003 [Abstract] [Google Scholar]
28. Chan SY, Zhang YY, Hemann C, Mahoney CE, Zweier JL, Loscalzo J. MicroRNA-210 controls mitochondrial metabolism during hypoxia by repressing the iron-sulfur cluster assembly proteins ISCU1/2. Cell Metab 10: 273–284, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
29. Chance B, Williams GR. The respiratory chain and oxidative phosphorylation. Adv Enzymol Relat Subj Biochem 17: 65–134, 1956 [Abstract] [Google Scholar]
30. Chandel N, Budinger GR, Kemp RA, Schumacker PT. Inhibition of cytochrome-c oxidase activity during prolonged hypoxia. Am J Physiol Lung Cell Mol Physiol 268: L918–L925, 1995 [Abstract] [Google Scholar]
31. Chandel NS, Budinger GR, Choe SH, Schumacker PT. Cellular respiration during hypoxia. Role of cytochrome oxidase as the oxygen sensor in hepatocytes. J Biol Chem 272: 18808–18816, 1997 [Abstract] [Google Scholar]
32. Chandel NS, Budinger GR, Schumacker PT. Molecular oxygen modulates cytochrome c oxidase function. J Biol Chem 271: 18672–18677, 1996. [Abstract] [Google Scholar]
33. Chen Z, Li Y, Zhang H, Huang P, Luthra R. Hypoxia-regulated microRNA-210 modulates mitochondrial function and decreases ISCU and COX10 expression. Oncogene 29: 4362–4368 [Abstract] [Google Scholar]
34. Chibalin AV, Ogimoto G, Pedemonte CH, Pressley TA, Katz AI, Feraille E, Berggren PO, Bertorello AM. Dopamine-induced endocytosis of Na+,K+-ATPase is initiated by phosphorylation of Ser-18 in the rat alpha subunit and Is responsible for the decreased activity in epithelial cells. J Biol Chem 274: 1920–1927, 1999 [Abstract] [Google Scholar]
35. Cleeter MW, Cooper JM, Darley-Usmar VM, Moncada S, Schapira AH. Reversible inhibition of cytochrome c oxidase, the terminal enzyme of the mitochondrial respiratory chain, by nitric oxide. Implications for neurodegenerative diseases. FEBS Lett 345: 50–54, 1994 [Abstract] [Google Scholar]
36. Clementi E, Brown GC, Foxwell N, Moncada S. On the mechanism by which vascular endothelial cells regulate their oxygen consumption. Proc Natl Acad Sci USA 96: 1559–1562, 1999 [Europe PMC free article] [Abstract] [Google Scholar]
37. Comellas AP, Dada LA, Lecuona E, Pesce LM, Chandel NS, Quesada N, Budinger GR, Strous GJ, Ciechanover A, Sznajder JI. Hypoxia-mediated degradation of Na,K-ATPase via mitochondrial reactive oxygen species and the ubiquitin-conjugating system. Circ Res 98: 1314–1322, 2006 [Abstract] [Google Scholar]
38. Dada LA, Chandel NS, Ridge KM, Pedemonte C, Bertorello AM, Sznajder JI. Hypoxia-induced endocytosis of Na,K-ATPase in alveolar epithelial cells is mediated by mitochondrial reactive oxygen species and PKC-zeta. J Clin Invest 111: 1057–1064, 2003 [Europe PMC free article] [Abstract] [Google Scholar]
39. Emerling BM, Weinberg F, Snyder C, Burgess Z, Mutlu GM, Viollet B, Budinger GR, Chandel NS. Hypoxic activation of AMPK is dependent on mitochondrial ROS but independent of an increase in AMP/ATP ratio. Free Radic Biol Med 46: 1386–1391, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
40. Enerback S. Human brown adipose tissue. Cell Metab 11: 248–252 [Abstract] [Google Scholar]
41. Epstein AC, Gleadle JM, McNeill LA, Hewitson KS, O'Rourke J, Mole DR, Mukherji M, Metzen E, Wilson MI, Dhanda A, Tian YM, Masson N, Hamilton DL, Jaakkola P, Barstead R, Hodgkin J, Maxwell PH, Pugh CW, Schofield CJ, Ratcliffe PJC. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell 107: 43–54, 2001. [Abstract] [Google Scholar]
42. Favaro E, Ramachandran A, McCormick R, Gee H, Blancher C, Crosby M, Devlin C, Blick C, Buffa F, Li JL, Vojnovic B, Pires das Neves R, Glazer P, Iborra F, Ivan M, Ragoussis Jand, Harris AL. MicroRNA-210 regulates mitochondrial free radical response to hypoxia and krebs cycle in cancer cells by targeting iron sulfur cluster protein ISCU. PLoS One 5: e10345. [Europe PMC free article] [Abstract] [Google Scholar]
43. Firth JD, Ebert BL, Ratcliffe PJ. Hypoxic regulation of lactate dehydrogenase A. Interaction between hypoxia-inducible factor 1 and cAMP response elements. J Biol Chem 270: 21021–21027, 1995 [Abstract] [Google Scholar]
44. Fukuda R, Zhang H, Kim JW, Shimoda L, Dang CV, Semenza GL. HIF-1 regulates cytochrome oxidase subunits to optimize efficiency of respiration in hypoxic cells. Cell 129: 111–122, 2007 [Abstract] [Google Scholar]
45. Gerald D, Berra E, Frapart YM, Chan DA, Giaccia AJ, Mansuy D, Pouyssegur J, Yaniv M, Mechta-Grigoriou F. JunD reduces tumor angiogenesis by protecting cells from oxidative stress. Cell 118: 781–794, 2004 [Abstract] [Google Scholar]
46. Gusarova GA, Dada LA, Kelly AM, Brodie C, Witters LA, Chandel NS, Sznajder JI. Alpha1-AMP-activated protein kinase regulates hypoxia-induced Na,K-ATPase endocytosis via direct phosphorylation of protein kinase C zeta. Mol Cell Biol 29: 3455–3464, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
47. Hamanaka RB, Chandel NS. Mitochondrial reactive oxygen species regulate cellular signaling and dictate biological outcomes. Trends Biochem Sci 35: 505–513 [Europe PMC free article] [Abstract] [Google Scholar]
48. Hardie DG. AMP-activated/SNF1 protein kinases: conserved guardians of cellular energy. Nat Rev Mol Cell Biol 8: 774–785, 2007 [Abstract] [Google Scholar]
49. Hawley SA, Boudeau J, Reid JL, Mustard KJ, Udd L, Makela TP, Alessi DR, Hardie DG. Complexes between the LKB1 tumor suppressor, STRAD alpha/beta and MO25 alpha/beta are upstream kinases in the AMP-activated protein kinase cascade. J Biol 2: 28, 2003 [Europe PMC free article] [Abstract] [Google Scholar]
50. Hawley SA, Pan DA, Mustard KJ, Ross L, Bain J, Edelman AM, Frenguelli BG, Hardie DG. Calmodulin-dependent protein kinase kinase-beta is an alternative upstream kinase for AMP-activated protein kinase. Cell Metab 2: 9–19, 2005 [Abstract] [Google Scholar]
51. Hay N, Sonenberg N. Upstream and downstream of mTOR. Genes Dev 18: 1926–1945, 2004 [Abstract] [Google Scholar]
52. Hochachka PW. Defense strategies against hypoxia and hypothermia. Science 231: 234–241, 1986 [Abstract] [Google Scholar]
53. Hochachka PW, Buck LT, Doll CJ, Land SC. Unifying theory of hypoxia tolerance: molecular/metabolic defense and rescue mechanisms for surviving oxygen lack. Proc Natl Acad Sci USA 93: 9493–9498, 1996 [Europe PMC free article] [Abstract] [Google Scholar]
54. Holness MJ, Sugden MC. Regulation of pyruvate dehydrogenase complex activity by reversible phosphorylation. Biochem Soc Trans 31: 1143–1151, 2003 [Abstract] [Google Scholar]
55. Hon WC, Wilson MI, Harlos K, Claridge TD, Schofield CJ, Pugh CW, Maxwell PH, Ratcliffe PJ, Stuart DI, Jones EY. Structural basis for the recognition of hydroxyproline in HIF-1 alpha by pVHL. Nature 417: 975–978, 2002 [Abstract] [Google Scholar]
56. Horman S, Browne G, Krause U, Patel J, Vertommen D, Bertrand L, Lavoinne A, Hue L, Proud C, Rider M. Activation of AMP-activated protein kinase leads to the phosphorylation of elongation factor 2 and an inhibition of protein synthesis. Curr Biol 12: 1419–1423, 2002 [Abstract] [Google Scholar]
57. Huang X, Ding L, Bennewith KL, Tong RT, Welford SM, Ang KK, Story M, Le QT, Giaccia AJ. Hypoxia-inducible mir-210 regulates normoxic gene expression involved in tumor initiation. Mol Cell 35: 856–867, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
58. Hudson CC, Liu M, Chiang GG, Otterness DM, Loomis DC, Kaper F, Giaccia AJ, Abraham RT. Regulation of hypoxia-inducible factor 1alpha expression and function by the mammalian target of rapamycin. Mol Cell Biol 22: 7004–7014, 2002 [Europe PMC free article] [Abstract] [Google Scholar]
59. Humar R, Kiefer FN, Berns H, Resink TJ, Battegay EJ. Hypoxia enhances vascular cell proliferation and angiogenesis in vitro via rapamycin (mTOR)-dependent signaling. FASEB J 16: 771–780, 2002 [Abstract] [Google Scholar]
60. Hyvarinen J, Hassinen IE, Sormunen R, Maki JM, Kivirikko KI, Koivunen P, Myllyharju J. Hearts of hypoxia-inducible factor prolyl 4-hydroxylase-2 hypomorphic mice show protection against acute ischemia-reperfusion injury. J Biol Chem 285: 13646–13657 [Europe PMC free article] [Abstract] [Google Scholar]
61. Ivan M, Kondo K, Yang H, Kim W, Valiando J, Ohh M, Salic A, Asara JM, Lane WS, Kaelin WG., Jr HIFalpha targeted for VHL-mediated destruction by proline hydroxylation: implications for O2 sensing. Science 292: 464–468, 2001 [Abstract] [Google Scholar]
62. Iyer NV, Kotch LE, Agani F, Leung SW, Laughner E, Wenger RH, Gassmann M, Gearhart JD, Lawler AM, Yu AY, Semenza GL. Cellular and developmental control of O2 homeostasis by hypoxia-inducible factor 1 alpha. Genes Dev 12: 149–162, 1998 [Europe PMC free article] [Abstract] [Google Scholar]
63. Jaakkola P, Mole DR, Tian YM, Wilson MI, Gielbert J, Gaskell SJ, Kriegsheim A, Hebestreit HF, Mukherji M, Schofield CJ, Maxwell PH, Pugh CW, Ratcliffe PJ. Targeting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science 292: 468–472, 2001 [Abstract] [Google Scholar]
64. Jones DP, Shan X, Park Y. Coordinated multisite regulation of cellular energy metabolism. Annu Rev Nutr 12: 327–343, 1992 [Abstract] [Google Scholar]
65. Jung F, Palmer LA, Zhou N, Johns RA. Hypoxic regulation of inducible nitric oxide synthase via hypoxia inducible factor-1 in cardiac myocytes. Circ Res 86: 319–325, 2000 [Abstract] [Google Scholar]
66. Kacser H, Burns JA. The control of flux. Symp Soc Exp Biol 27: 65–104, 1973 [Abstract] [Google Scholar]
67. Kaelin WG, Jr, Ratcliffe PJ. Oxygen sensing by metazoans: the central role of the HIF hydroxylase pathway. Mol Cell 30: 393–402, 2008 [Abstract] [Google Scholar]
68. Kahn BB, Alquier T, Carling D, Hardie DG. AMP-activated protein kinase: ancient energy gauge provides clues to modern understanding of metabolism. Cell Metab 1: 15–25, 2005 [Abstract] [Google Scholar]
69. Kaplan JH. Biochemistry of Na,K-ATPase. Annu Rev Biochem 71: 511–535, 2002 [Abstract] [Google Scholar]
70. Kido M, Du L, Sullivan CC, Li X, Deutsch R, Jamieson SW, Thistlethwaite PA. Hypoxia-inducible factor 1-alpha reduces infarction and attenuates progression of cardiac dysfunction after myocardial infarction in the mouse. J Am Coll Cardiol 46: 2116–2124, 2005 [Abstract] [Google Scholar]
71. Kim JW, Tchernyshyov I, Semenza GL, Dang CV. HIF-1-mediated expression of pyruvate dehydrogenase kinase: a metabolic switch required for cellular adaptation to hypoxia. Cell Metab 3: 177–185, 2006 [Abstract] [Google Scholar]
72. Koritzinsky M, Magagnin MG, vanden Beucken T, Seigneuric R, Savelkouls K, Dostie J, Pyronnet S, Kaufman RJ, Weppler SA, Voncken JW, Lambin P, Koumenis C, Sonenberg N, Wouters BG. Gene expression during acute and prolonged hypoxia is regulated by distinct mechanisms of translational control. EMBO J 25: 1114–1125, 2006 [Europe PMC free article] [Abstract] [Google Scholar]
73. Koumenis C, Naczki C, Koritzinsky M, Rastani S, Diehl A, Sonenberg N, Koromilas A, Wouters BG. Regulation of protein synthesis by hypoxia via activation of the endoplasmic reticulum kinase PERK and phosphorylation of the translation initiation factor eIF2alpha. Mol Cell Biol 22: 7405–7416, 2002 [Europe PMC free article] [Abstract] [Google Scholar]
74. Kulshreshtha R, Ferracin M, Wojcik SE, Garzon R, Alder H, Agosto-Perez FJ, Davuluri R, Liu CG, Croce CM, Negrini M, Calin GA, Ivan M. A microRNA signature of hypoxia. Mol Cell Biol 27: 1859–1867, 2007 [Europe PMC free article] [Abstract] [Google Scholar]
75. Lee SC, Downey HF. Downregulation of oxygen demand in isoprenaline stimulated canine myocardium. Cardiovasc Res 27: 1542–1550, 1993 [Abstract] [Google Scholar]
76. Lee VY, McClintock DS, Santore MT, Budinger GR, Chandel NS. Hypoxia sensitizes cells to nitric oxide-induced apoptosis. J Biol Chem 277: 16067–16074, 2002 [Abstract] [Google Scholar]
77. Liu L, Cash TP, Jones RG, Keith B, Thompson CB, Simon MC. Hypoxia-induced energy stress regulates mRNA translation and cell growth. Mol Cell 21: 521–531, 2006 [Europe PMC free article] [Abstract] [Google Scholar]
78. Liu L, Simon MC. Regulation of transcription and translation by hypoxia. Cancer Biol Ther 3: 492–497, 2004 [Abstract] [Google Scholar]
79. Liu L, Wise DR, Diehl JA, Simon MC. Hypoxic reactive oxygen species regulate the integrated stress response and cell survival. J Biol Chem 283: 31153–31162, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
80. Lizasoain I, Moro MA, Knowles RG, Darley-Usmar V, Moncada S. Nitric oxide and peroxynitrite exert distinct effects on mitochondrial respiration which are differentially blocked by glutathione or glucose. Biochem J 314: 877–880, 1996 [Europe PMC free article] [Abstract] [Google Scholar]
81. Mairbaurl H, Wodopia R, Eckes S, Schulz S, Bartsch P. Impairment of cation transport in A549 cells and rat alveolar epithelial cells by hypoxia. Am J Physiol Lung Cell Mol Physiol 273: L797–L806, 1997 [Abstract] [Google Scholar]
82. Mason SD, Rundqvist H, Papandreou I, Duh R, McNulty WJ, Howlett RA, Olfert IM, Sundberg CJ, Denko NC, Poellinger L, Johnson RS. HIF-1α in endurance training: suppression of oxidative metabolism. Am J Physiol Regul Integr Comp Physiol 293: R2059–R2069, 2007 [Abstract] [Google Scholar]
83. Masson N, Willam C, Maxwell PH, Pugh CW, Ratcliffe PJ. Independent function of two destruction domains in hypoxia-inducible factor-alpha chains activated by prolyl hydroxylation. EMBO J 20: 5197–5206, 2001 [Europe PMC free article] [Abstract] [Google Scholar]
84. Maxwell PH, Wiesener MS, Chang GW, Clifford SC, Vaux EC, Cockman ME, Wykoff CC, Pugh CW, Maher ER, Ratcliffe PJ. The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature 399: 271–275, 1999 [Abstract] [Google Scholar]
85. Melillo G, Taylor LS, Brooks A, Musso T, Cox GW, Varesio L. Functional requirement of the hypoxia-responsive element in the activation of the inducible nitric oxide synthase promoter by the iron chelator desferrioxamine. J Biol Chem 272: 12236–12243, 1997 [Abstract] [Google Scholar]
86. Milligan LP, McBride BW. Energy costs of ion pumping by animal tissues. J Nutr 115: 1374–1382, 1985 [Abstract] [Google Scholar]
87. Natarajan R, Salloum FN, Fisher BJ, Kukreja RC, Fowler AA., 3rd Hypoxia inducible factor-1 upregulates adiponectin in diabetic mouse hearts and attenuates post-ischemic injury. J Cardiovasc Pharmacol 51: 178–187, 2008 [Abstract] [Google Scholar]
88. Pantely GA, Bristow JD. Hibernating myocardium: a hypometabolic state for energy conservation. Basic Res Cardiol 90: 23–25, 1995 [Abstract] [Google Scholar]
89. Papandreou I, Cairns RA, Fontana L, Lim AL, Denko NC. HIF-1 mediates adaptation to hypoxia by actively downregulating mitochondrial oxygen consumption. Cell Metab 3: 187–197, 2006 [Abstract] [Google Scholar]
90. Planes C, Friedlander G, Loiseau A, Amiel C, Clerici C. Inhibition of Na-K-ATPase activity after prolonged hypoxia in an alveolar epithelial cell line. Am J Physiol Lung Cell Mol Physiol 271: L70–L78, 1996 [Abstract] [Google Scholar]
91. Puissegur MP, Mazure NM, Bertero T, Pradelli L, Grosso S, Robbe-Sermesant K, Maurin T, Lebrigand K, Cardinaud B, Hofman V, Fourre S, Magnone V, Ricci JE, Pouyssegur J, Gounon P, Hofman P, Barbry P, Mari B. miR-210 is overexpressed in late stages of lung cancer and mediates mitochondrial alterations associated with modulation of HIF-1 activity. Cell Death Differ. 2010. October 1 [Epub ahead of print.] [Europe PMC free article] [Abstract] [Google Scholar]
92. Ridge KM, Dada L, Lecuona E, Bertorello AM, Katz AI, Mochly-Rosen D, Sznajder JI. Dopamine-induced exocytosis of Na,K-ATPase is dependent on activation of protein kinase C-epsilon and -delta. Mol Biol Cell 13: 1381–1389, 2002 [Europe PMC free article] [Abstract] [Google Scholar]
93. Saraste M. Oxidative phosphorylation at the fin de siecle. Science 283: 1488–1493, 1999 [Abstract] [Google Scholar]
94. Schneider M, Van Geyte K, Fraisl P, Kiss J, Aragones J, Mazzone M, Mairbaurl H, De Bock K, Jeoung NH, Mollenhauer M, Georgiadou M, Bishop T, Roncal C, Sutherland A, Jordan B, Gallez B, Weitz J, Harris RA, Maxwell P, Baes M, Ratcliffe P, Carmeliet P. Loss or silencing of the PHD1 prolyl hydroxylase protects livers of mice against ischemia/reperfusion injury. Gastroenterology 138: 1143–1154, e1141–e1142 [Abstract] [Google Scholar]
95. Schroder M, Kaufman RJ. The mammalian unfolded protein response. Annu Rev Biochem 74: 739–789, 2005 [Abstract] [Google Scholar]
96. Schumacker PT, Chandel N, Agusti AG. Oxygen conformance of cellular respiration in hepatocytes. Am J Physiol Lung Cell Mol Physiol 265: L395–L402, 1993 [Abstract] [Google Scholar]
97. Semenza GL. Life with oxygen. Science 318: 62–64, 2007 [Abstract] [Google Scholar]
98. Semenza GL. Oxygen-dependent regulation of mitochondrial respiration by hypoxia-inducible factor 1. Biochem J 405: 1–9, 2007. [Abstract] [Google Scholar]
99. Sengupta S, Peterson TR, Sabatini DM. Regulation of the mTOR complex 1 pathway by nutrients, growth factors, and stress. Mol Cell 40 310–322 [Europe PMC free article] [Abstract] [Google Scholar]
100. Shaw RJ. LKB1 and AMP-activated protein kinase control of mTOR signalling and growth. Acta Physiol (Oxf) 196: 65–80, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
101. Shaw RJ, Kosmatka M, Bardeesy N, Hurley RL, Witters LA, De Pinho 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 USA 101: 3329–3335, 2004 [Europe PMC free article] [Abstract] [Google Scholar]
102. Shim H, Chun YS, Lewis BC, Dang CV. A unique glucose-dependent apoptotic pathway induced by c-Myc. Proc Natl Acad Sci USA 95: 1511–1516, 1998 [Europe PMC free article] [Abstract] [Google Scholar]
103. Van der Heiden MG, Cantley LC, Thompson CB. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science 324: 1029–1033, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
104. Wang GL, Semenza GL. Purification and characterization of hypoxia-inducible factor 1. J Biol Chem 270: 1230–1237, 1995 [Abstract] [Google Scholar]
105. Ward JP, Snetkov VA, Aaronson PI. Calcium, mitochondria and oxygen sensing in the pulmonary circulation. Cell Calcium 36: 209–220, 2004 [Abstract] [Google Scholar]
106. Waypa GB, Marks JD, Mack MM, Boriboun C, Mungai PT, Schumacker PT. Mitochondrial reactive oxygen species trigger calcium increases during hypoxia in pulmonary arterial myocytes. Circ Res 91: 719–726, 2002 [Abstract] [Google Scholar]
107. Wellen KE, Thompson CB. Cellular metabolic stress: considering how cells respond to nutrient excess. Mol Cell 40: 323–332 [Europe PMC free article] [Abstract] [Google Scholar]
108. Wilson DF, Erecinska M, Stubbs M, Lindsay JG, Owen CS. Cellular control of mitochondrial respiration. Adv Exp Med Biol 75: 137–144, 1976 [Abstract] [Google Scholar]
109. Wilson DF, Rumsey WL, Green TJ, Vanderkooi JM. The oxygen dependence of mitochondrial oxidative phosphorylation measured by a new optical method for measuring oxygen concentration. J Biol Chem 263: 2712–2718, 1988 [Abstract] [Google Scholar]
110. Wodopia R, Ko HS, Billian J, Wiesner R, Bartsch P, Mairbaurl H. Hypoxia decreases proteins involved in epithelial electrolyte transport in A549 cells and rat lung. Am J Physiol Lung Cell Mol Physiol 279: L1110–L1119, 2000 [Abstract] [Google Scholar]
111. 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 13: 2004–2008, 2003 [Abstract] [Google Scholar]
112. Wu IC, Ohsawa I, Fuku N, Tanaka M. Metabolic analysis of 13C-labeled pyruvate for noninvasive assessment of mitochondrial function. Ann NY Acad Sci 1201: 111–120 [Abstract] [Google Scholar]
113. Zhou G, Dada LA, Chandel NS, Iwai K, Lecuona E, Ciechanover A, Sznajder JI. Hypoxia-mediated Na-K-ATPase degradation requires von Hippel Lindau protein. FASEB J 22: 1335–1342, 2008 [Abstract] [Google Scholar]

Articles from American Journal of Physiology - Cell Physiology are provided here courtesy of American Physiological Society

Citations & impact 


Impact metrics

Jump to Citations

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/6105422
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/6105422

Article citations


Go to all (226) article citations

Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.

Funding 


Funders who supported this work.

NCI NIH HHS (2)