Oxygen is essential to the life of all aerobic organisms. Virtually every cell type is able to sense a limited oxygen supply (hypoxia) and specifically to induce a set of oxygen-regulated genes. This review summarizes current concepts of mammalian oxygen-sensing and signal-transduction pathways. Since the discovery of the hypoxia-inducible factors (HIFs), a great deal of progress has been made in our comprehension of how hypoxia induces the expression of oxygen-regulated genes. The α subunit of the heterodimeric transcription factors HIF-1, 2 and 3 is unstable under normoxia but is rapidly stabilized upon exposure to hypoxic conditions. Following heterodimerization with the constitutively expressed β subunit, HIFs activate the transcription of an increasing number of genes involved in maintaining oxygen homeostasis at the cellular, local and systemic levels.

An adequate supply of oxygen is essential to all higher organisms because it serves as the terminal electron acceptor in mitochondrial oxidative phosphorylation and because several enzymatic processes require molecular oxygen as a substrate. Even slight reductions in normal oxygen concentrations (hypoxia) can cause the induction of specific genes involved in mammalian oxygen homeostasis such as erythropoietin or vascular endothelial growth factor (VEGF). Investigations of such hypoxia-inducible genes performed in many different cultured cell lines suggest that every mammalian (perhaps even every vertebrate and insect) cell possesses one or several oxygen-sensing mechanism(s), i.e. a molecular oxygen sensor. Certain cell types, which can be found in carotid bodies, airway neuroepithelial bodies and pulmonary arteries, have differentiated into specialized oxygen-sensing cells. However, the mammalian cellular oxygen sensor is not known. The beautifully elaborated principles of oxygen sensing in bacteria and yeast (for a review, see Bunn and Poyton, 1996) have still to find their counterparts in higher organisms. In the nitrogen-fixing bacterium Bradyrhizobium japonicum, for instance, the oxygen sensor is a histidine kinase termed FixL which, under anaerobic conditions, phosphorylates and thereby activates the transcription factor FixJ. Interestingly, the oxygen-sensing domain of FixL is a so-called PAS domain (see below), a versatile protein fold resembling a left-handed glove that encloses a haem cofactor (Gong et al., 1998). When oxygen is bound to ferrous (FeII) haem, this domain is in the flattened ‘off’ state, resulting in a change in the protein conformation and inactivation of the kinase domain.

The haem oxygen sensor hypothesis

So far, virtually all proteins capable of binding molecular oxygen contain iron, usually in the centre of a haem moiety. Thus, it came as no surprise when, in 1988, Bunn and co-workers suggested that the mammalian oxygen sensor could be a haem protein (Goldberg et al., 1988). Apart from hypoxia, cobalt (and to a lesser extent nickel and manganese) salts were capable of inducing erythropoietin gene expression with no additional effects under hypoxic conditions, indicating that ferrous haem iron was replaced by the non-oxygen-binding cations. According to the model, this locks the oxygen sensor in the deoxy conformation. Moreover, carbon monoxide inhibited hypoxic but not cobalt-mediated activation of erythropoietin gene expression, in line with the known properties of carbon monoxide binding to haemoglobin. Blockers of haem synthesis abolished both hypoxic and cobalt-dependent erythropoietin induction. Further evidence for the haem hypothesis was provided by the finding that iron chelators, such as desferrioxamine, are also capable of mimicking hypoxia (Ho and Bunn, 1996; Wang and Semenza, 1993b).

Although the basic findings that led to the hypothesis of a haem oxygen sensor responsible for hypoxic erythropoietin induction have been confirmed for many other hypoxia-inducible genes, several discrepancies remain. For example, differential effects of the flavoprotein inhibitor diphenylene iodonium on hypoxia-induced versus cobalt-induced gene expression have been reported. Ratcliffe and co-workers found that 50 nmol l−1 and 1 μmol l−1 diphenylene iodonium inhibited hypoxic but not cobalt- or iron-chelator-dependent induction of several oxygen-regulated genes (Gleadle et al., 1995). In striking contrast, Goldwasser et al. (1995) demonstrated a 50 % inhibition of cobalt chloride but not of hypoxic erythropoietin induction by 30 nmol l−1 diphenylene iodonium. Schumacker and co-workers reported that diphenylene iodonium and two other mitochondrial inhibitors (rotenone and myxothiazol) blocked hypoxic but not cobalt-dependent induction of erythropoietin mRNA (Chandel et al., 1998). Ratcliffe and co-workers (Gleadle et al., 1995) and Semenza and co-workers (Agani and Semenza, 1998) also showed that the flavoprotein cytochrome P450 reductase inhibitor mersalyl inhibited hypoxia-, cobalt-chloride- and desferrioxamine-dependent induction of the erythropoietin gene. However, mersalyl induced the hypoxia-inducible factor-1α (HIF-1α, see below) and the vascular endothelial growth factor (VEGF) gene (Agani and Semenza, 1998). Mersalyl also induced a reporter gene under the control of transferrin-derived hypoxia-responsive elements (Rolfs et al., 1997) under normoxia without any further effects under hypoxia (R. Wanner and R. H. Wenger, in preparation). Finally, following careful titration of oxygen and cobalt chloride concentrations (taking into consideration that the optimal concentration of the toxic agent cobalt chloride is itself a function of the oxygen concentration), we found that there might well be additive effects on erythropoietin induction when both agents were applied simultaneously (Wenger et al., 1998b).

In summary, these discrepancies suggest either that hypoxia, cobalt and desferrioxamine interact with different steps along the oxygen signal-transduction pathway or that they each affect additional pathways that cross-talk with the oxygen pathway. These contradictory reports also reflect the difficulties of using pleiotropic agents to track down the oxygen sensor to its molecular entity. Different effects of these agents depending on concentration, target genes and cell types have to be considered. Nevertheless, the haem hypothesis is still the best model available. But what kind of haem protein is it?

Many candidates – no election

Because there is no assay available to identify directly the oxygen sensor among the hundreds of known oxygen-binding proteins, many different candidate proteins have been suggested in the past to play this role as a sensor in addition to their known or proposed functions. The following section will provide some examples and discuss briefly the pros and cons for their putative role in oxygen sensing.

One of the most often cited putative haem oxygen-sensing molecules is the cytochrome b558/NADPH oxidase complex that generates superoxide in the plasma membrane of phagocytes and B lymphocytes. The subunits of this complex have also been found in cells of non-immunological origin, including hepatoma cell lines (Görlach et al., 1993, 1994), small cell lung carcinoma (Wang et al., 1996) and pheochromocytoma PC12 cells (Hohler et al., 1999), and in oxygen-sensing cells of the carotid body (the arterial chemoreceptor; Acker et al., 1989; Cross et al., 1990; Kummer and Acker, 1995; Lahiri et al., 1999) and pulmonary neuroepithelial bodies (the airway chemoreceptors; Wang et al., 1996; Youngson et al., 1993, 1997). Therefore, Acker and co-workers put forward the hypothesis that the b558 complex may serve as an oxygen sensor (for a review, see Acker, 1994). However, we have demonstrated normal oxygen-regulated gene expression in cell lines derived from patients suffering from chronic granulomatous disease, an inherited disease in which one of the subunits of the b558/NADPH oxidase complex is defective (Wenger et al., 1996). Our conclusion that this complex is not likely to be the ‘universal’ oxygen sensor has recently been confirmed in knock-out mice deficient in the gp91phox subunit of the b558/NADPH oxidase complex (Archer et al., 1999). Despite a marked reduction in superoxide production, hypoxia caused similar levels of pulmonary vasoconstriction and inhibition of whole-cell K+ current in both knock-out and wild-type mice.

Another apparently promising candidate oxygen sensor is the principal oxygen-consuming organelle of the cell, the mitochondrion itself. Because the respiratory electron transport chain blocker potassium cyanide cannot induce erythropoietin gene expression, mitochondria were soon ruled out as the oxygen sensor (Goldberg et al., 1988; Tan and Ratcliffe, 1991). Nevertheless, Lahiri and co-workers proposed that mitochondrial cytochrome a3 (Lahiri et al., 1995; Wilson et al., 1994) and Schumacker and co-workers suggested that mitochondrial cytochrome c oxidase (Chandel et al., 1997) could serve as the oxygen sensor. Interestingly, depleting cells of mitochondrial DNA (μ0 cells) and treatment with several mitochondrial inhibitors blocked hypoxic induction of oxygen-regulated genes (Chandel et al., 1998). However, because potassium cyanide does not interfere with hypoxic induction of HIF-1α (Jiang et al., 1996), the role of mitochondria in oxygen sensing remains unclear.

Various enzymatic processes that result in the production of different second-messenger molecules are catalyzed by haem proteins. For instance, the guanylate cyclase and cytochrome P450 of the 4A family, producing cyclic GMP (cGMP) and 20-hydroxyeicosatetraenoic acid (20-HETE), respectively, have been proposed as oxygen sensors (Harder et al., 1996; Taylor et al., 1998). Cytochrome P450 (ant)agonists have also been shown to modulate erythropoietin expression (Fandrey et al., 1990), but the involvement of cytochrome P450 could not be confirmed by spectrophotometric analysis (Ehleben et al., 1998).

What next

As outlined above, the ‘comparative approach’ to the identification of the oxygen sensor is biased by our limited knowledge of the function and the number of candidate haem proteins. There may be many more candidates since a haem protein cannot be recognized easily from its primary amino acid sequence. What would be an unbiased ‘empirical approach’ to this problem? Oxygen binding is not a sufficient criterion for the sensor; it must connect an oxygen-dependent enzymatic activity (many of which are already known) to the regulation of hypoxia-inducible genes. Therefore, only the inactivation of the (unknown) sensor and an analysis of the consequences on gene activation can serve to identify it unambiguously. Ratcliffe and co-workers used hypoxia-response elements to direct the expression of reporter genes. Following random mutagenesis, cells were selected that had lost the ability for oxygen-regulated gene expression. This approach has been validated by the identification of an HIF-1α inactivation mutation (Wood et al., 1998). The oxygen sensor, however, remained unidentified.

The reactive oxygen species dilemma

Reactive oxygen species (ROS) such as superoxide and peroxide are known to serve as signal transducers in several systems (for a review, see Suzuki et al., 1997). Because oxygen is the main component of these molecules, it seems obvious that their concentration depends upon the environmental oxygen concentration. Fandrey, Jelkmann and co-workers reported a close relationship between oxygen and hydrogen peroxide concentrations and rates of erythropoietin production in hepatoma cells (Fandrey et al., 1994). Hydrogen peroxide sequestration could mimic hypoxia (Canbolat et al., 1998) and addition of hydrogen peroxide inhibited hypoxic induction of erythropoietin (Fandrey et al., 1994). This inhibition could be reversed by cobalt chloride and iron chelation (Fandrey et al., 1997). These findings have been confirmed for the oxygen-regulated genes encoding tyrosine hydroxylase (Kroll and Czyzyk-Krzeska, 1998) and several enzymes involved in glucose metabolism (Kietzmann et al., 1996, 1997). This model is also compatible with the idea that the thiol reducing agent N-(2-mercaptopropionyl)-glycine and the oxygen radical scavenger 2-acetamidoacrylic acid, but not many other antioxidants, could mimic hypoxia by inducing HIF-1α under normoxic conditions (Chilov et al., 1999; Salceda and Caro, 1997).

Enzymes generating ROS, such as the cytochrome b558/ NADPH oxidase complex (see above), have been proposed as the source for superoxide, which is rapidly converted into peroxide by superoxide dismutase. Acker and co-workers suggested a localized Fenton reaction in perinuclear cytosolic granules containing high iron concentrations (Kietzmann et al., 1998; Porwol et al., 1998). The Fenton reaction describes the non-enzymatic reduction of hydrogen peroxide to a hydroxyl anion and the highly reactive hydroxyl radical by oxidation of the ferrous iron. This model would separate the sites of oxygen sensing and the effects of cobalt/iron chelation, and it would explain the differences observed in the function of these agents. According to this hypothesis, normoxic hydroxyl radical concentrations would inactivate HIF-1α, which is relieved from suppression following a decrease in radical concentrations under hypoxic conditions. A prototype factor for such a mechanism is the iron regulatory protein 2, which is oxidized in an oxygen-dependent manner, ubiquitinated and degraded by the proteasomes in iron-replete cells (Iwai et al., 1998).

In striking contrast to these findings, Schumacker and co-workers reported that cardiomyocytes increase their production of ROS following exposure to hypoxia (Duranteau et al., 1998; Vanden Hoek et al., 1998). Antioxidants inhibited this increase in levels of ROS and also blocked hypoxic inhibition of cardiomyocyte contraction. The mitochondrial cytochrome c oxidase was suggested as the oxygen sensor that enhanced ROS production in hypoxic cardiomyocytes (Budinger et al., 1998) and hepatocytes (Chandel et al., 1997). Support for this model was provided by using hepatoma cells depleted of functional mitochondria (μ0 cells), which showed no induction of ROS under hypoxia and which failed to induce oxygen-regulated gene expression (Chandel et al., 1998).

Taken together, there is no common consensus about whether ROS concentrations decrease or increase following exposure to acute hypoxia. Further experiments determining the kinetics and dependence on oxygen concentration of ROS production will be required to obtain a more definitive picture.

Alternative or no second messengers

Apart from ROS, other metabolic processes occurring during hypoxia could potentially serve to sense and signal the cellular oxygenation state. A decrease in pH as a consequence of lactate production by anaerobic glycolysis might be responsible for hypoxic gene induction. The hypoxia-induced increase in ferritin production, for instance, might be regulated by iron release from ferritin as a result of acidification of the cytoplasm (Qi et al., 1995). Similarly, hypoxic induction of the p53 tumour suppressor protein has been reported to be dependent on acidosis (Schmaltz et al., 1998). However, a significant decrease in pH requires prolonged or severe hypoxia (anoxia) rather than the short periods of hypoxia sufficient to induce HIF-1α. Indeed, we have observed no effects of altered pH on oxygen-regulated gene expression (A. Rolfs and R. H. Wenger, unpublished observations). In general, it is difficult to understand how signal transducers such as ROS, redox state, pH, ATP, cAMP/cGMP or adenosine could trigger the ubiquitous, sensitive, specific and very rapid hypoxic induction of gene expression. A specific sensor that signals directly to HIF-1 and/or its degradation machinery seems to be more appropriate to fulfil these tasks.

The hypoxia-inducible factors

Compared with the relative lack of information on the oxygen sensor and its signal-transduction pathway, much more is known about the target of these putative signalling mechanisms. In 1995, Semenza and co-workers discovered HIF-1 on the basis of its ability to bind to a hypoxia-response element (HRE) in the 3′ flanking region of the erythropoietin gene (Wang et al., 1995a). HIF-1 is a ubiquitously and constitutively expressed heterodimeric transcription factor composed of an α subunit unstable under normoxia and a common β subunit, the latter being shared by other transcription factors (for reviews, see Bunn and Poyton, 1996; Semenza, 1998; Wenger and Gassmann, 1997). HIF-1β had previously been identified as the heterodimerization partner of the dioxin receptor/aryl hydrocarbon receptor (AhR) and was hence called AhR nuclear translocator (ARNT, see Hoffman et al., 1991). HIF-1α heterodimerization with ARNT in the nucleus is required for DNA binding and transactivation (Gassmann et al., 1997; Gradin et al., 1996; Kallio et al., 1997; Salceda et al., 1996; Wood et al., 1996) but not for translocation into the nucleus (Chilov et al., 1999; Kallio et al., 1998).

The most intriguing feature of HIF-1α is a so-called oxygen-dependent degradation domain (ODD) that can confer hypoxic stabilization to HIF-1α and to various fusion partner proteins (Huang et al., 1998; Pugh et al., 1997; Srinivas et al., 1999a). Under normoxic conditions, this domain seems to receive a signal that primes the HIF-1α protein for ubiquitination and subsequent degradation in proteasomes. In fact, HIF-1α protein is usually undetectable under normoxic conditions. Proteasomal inhibitors or mutation of the ubiquitin-activating enzyme E1 stabilize the HIF-1α protein but fail to induce oxygen-regulated gene expression (Huang et al., 1998; Kallio et al., 1999; Salceda and Caro, 1997). Normoxic HIF α subunit degradation involves the von-Hippel–Lindau tumour suppressor protein (pVHL) which recognizes and links HIF α subunits, via a complex composed of several incompletely characterized proteins, to the ubiquitination machinery (Maxwell et al., 1999). Cells deficient in functional pVHL show constitutively high HIF α subunit levels and expression of many oxygen-regulated genes, leading to haemangioblastoma tumour formation.

The signal that primes the ODD domain for degradation is not known. Reduction–oxidation (redox) reactions might be involved in normoxic HIF-1α degradation, as is the case for other proteins such as iron regulatory protein 2 (Iwai et al., 1998). Certain antioxidants and suppression of the Fenton reaction by iron-chelation can induce HIF-1α protein as well as HIF-1-dependent reporter gene activation (Huang et al., 1996; Salceda and Caro, 1997). However, oxidation of HIF-1α has not been demonstrated so far. Thus, it is not known whether this is a specific mechanism linking HIF-1α to the oxygen sensor or whether the experimental data obtained are the result of nonspecific processes and/or cross-talk with other signal-transduction pathways. It is an intriguing possibility that the autonomous ODD domain of HIF-1α might sense oxygen directly. This hypothesis would predict an iron-containing oxygen-binding domain, either as a haem protein or as an iron–sulphur cluster. However, we (P. Spielmann, F. Parpan and R. H. Wenger, unpublished observations) and others (Srinivas et al., 1999b) could not detect iron in bacterially expressed fragments of HIF-1α that contain the ODD domain.

Although phosphorylation seems to play an important role in the activation of HIF-1α (Salceda et al., 1997; Wang et al., 1995b), replacement of potential phosphoacceptor sites in the critical region of the ODD domain had no effect on its activity (Pugh et al., 1997; Srinivas et al., 1999a). Two recent reports showed that the p42/p44 mitogen-activated protein kinases (MAPKs) can phosphorylate the HIF α subunits, giving it electrophoretic migration properties similar to those of the hypoxically modified protein (Conrad et al., 1999; Richard et al., 1999). However, the phosphorylated amino acid residues have not been identified. MAPK-dependent phosphorylation enhances the transcriptional activity of HIFs, but in non-excitable cells the p42/p44 MAPKs are not activated by hypoxia (Richard et al., 1999). In excitable cells, hypoxia might indirectly activate MAPKs, but HIF α protein stabilization is not related to MAPK-dependent phosphorylation (Conrad et al., 1999). Thus, p42/p44 MAPKs do not account for the most critical step in HIF α subunit activation: the stabilization of the ODD domain.

HIF-1α contains two transactivation (TA) domains, one overlapping with the ODD domain and the second at the C terminus (Huang et al., 1998; Jiang et al., 1997; Li et al., 1996; Pugh et al., 1997). Apart from conferring protein stabilization, these transactivation domains might be further activated under hypoxic conditions (Kallio et al., 1998), for example via p42/p44 MAPK-dependent phosphorylation (Conrad et al., 1999; Richard et al., 1999) or redox-dependent processes (Huang et al., 1996; Salceda and Caro, 1997). A reductive environment provided by the redox factor Ref-1 enhances the recruitment of the transcriptional coactivators SRC-1, TIF2 and CBP/p300 (Carrero et al., 2000). Fujii-Kuriyama and co-workers identified a redox-sensitive conserved cysteine residue, mutation of which inhibited the transactivation function (Ema et al., 1999). These authors proposed a model in which reduced thioredoxin translocates into the nucleus of hypoxic cells and transfers the redox signal to Ref-1 which, in turn, interacts with this cysteine. Indeed, overexpression of thioredoxin/Ref-1 has been shown previously to amplify the hypoxic signal (Huang et al., 1996). However, the functional importance of this cysteine residue could not be confirmed in other mutation experiments (O’Rourke et al., 1999), and we did not observe nuclear translocation of thioredoxin under hypoxia (U. Alt and R. H. Wenger, unpublished observations). Therefore, the specificity of thioredoxin/Ref-1 in maintaining HIF-1α in a reduced state as part of the hypoxia-signalling mechanism awaits further clarification.

Apart from the ubiquitously expressed HIF-1α, two other members of this family, HIF-2α (Ema et al., 1997; Flamme et al., 1997; Hogenesch et al., 1997; Tian et al., 1997) and HIF-3α (Gu et al., 1998), have been identified which show a more restricted expression pattern. Functional comparison between HIF-1α and HIF-2α in vitro revealed many similarities concerning genomic organization, modular protein structure, hypoxic protein stabilization, heterodimerization, DNA binding and the transactivation function of reporter genes (Ema et al., 1999; Hogenesch et al., 1998; Maxwell et al., 1999; O’Rourke et al., 1999; Wenger et al., 1997; Wiesener et al., 1998). In vivo gene targeting (knock-out) experiments in mice showed that HIF-1α is required for mesenchymal cell survival during embryonic development: knock-out animals die around midgestation, showing cardiovascular malformations and open neural tube defects (Iyer et al., 1998; Kotch et al., 1999; Ryan et al., 1998). Mice containing only one mutant HIF-1α allele develop normally but show impaired physiological responses to chronic hypoxia (Yu et al., 1999). Interestingly, in HIF-2α knock-out embryos, the organ of Zuckerkandl (the major catecholamine source during development) does not produce sufficient amounts of catecholamines, leading to embryonic lethality presumably because of bradycardia (Tian et al., 1998). The expression pattern and functional role of HIF-3α have still to be elucidated.

The PAS family of transcription factors and sensors

Another interesting feature of the HIF α subunits is the presence of PAS domains which can also be found in mammalian ARNT1, 2 and 3 and in their corresponding heterodimerization partners (HIF-1α, 2α, 3α, AhR1, 2 and CLOCK) as well as in several other transcription factors including Per1, 2, Sim1, 2 and NPAS1, 2 and 3 (for a review, see Taylor and Zhulin, 1999). PAS is an acronym standing for the first family members known: Per, AhR/ARNT and Sim. Most of these PAS proteins have been cloned from mammalian and fish species and from Drosophila, where they play important roles in various aspects of development, oxygen homeostasis and circadian rhythm.

Most of the PAS domain proteins described to date have been found in the Bacteria and Archea kingdoms, where they often serve as sensors of oxygen levels, redox state or light levels. Intriguingly, the haem-containing pocket of the FixL oxygen sensor is a PAS domain (Gong et al., 1998), and it is tempting to speculate that the HIF-1α PAS domain displays similar features. However, haem binding to the PAS domain has never been detected, and it is unlikely that HIF-1α functions by itself as an oxygen sensor. In Eukarya, the PAS domain has apparently evolved to a heterodimerization interface that is usually preceded by a basic helix–loop–helix DNA-binding and heterodimerization domain. Whether functionally altered or not, it is astonishing that this domain is conserved in proteins of higher organisms that are functionally related to their predecessors in more primitive species.

HIF-1-mediated gene activation

Once activated by hypoxia, HIF-1 binds to the consensus HIF-1 DNA binding site (HBS) A/(G)CGTG present in the hypoxia-response elements (HREs) of many oxygen-regulated genes (for a review, see Wenger and Gassmann, 1997). In the erythropoietin gene, this site can also be constitutively bound by the transcription factors activating transcription factor-1 (ATF-1) and cAMP-responsive element-binding-1 (CREB-1) (Kvietikova et al., 1995, 1997). Normoxic occupancy of the HBS (Hu et al., 1997) might protect the CpG dinucleotide present in the HBS from becoming methylation-inactivated (Wenger et al., 1998a). For transactivation, HIF-1 recruits via the two transactivation domains of HIF-1α as well as the C-terminal transactivation domain of ARNT transcriptional coactivators including the histone acetyltransferases CBP/p300 (Arany et al., 1996; Ebert and Bunn, 1998; Ema et al., 1999; Kallio et al., 1998) and SRC-1 and TIF2 (Carrero et al., 2000). It should be noted that an HBS is necessary, but not sufficient, for efficient gene activation. Binding of other transcription factors in the vicinity of HIF-1 or multimerization of HBSs is usually observed. In the erythropoietin HRE, for example, an unknown factor binding to a CACA box plus the hepatocyte nuclear factor-4 are required to form a functional transcriptional complex (Galson et al., 1995). Other examples include cooperation with the ATF-1/CREB-1 factor in the lactate dehydrogenase A gene (Ebert and Bunn, 1998; Firth et al., 1995) or with activator protein-1 (AP-1) binding factors in the VEGF gene (Damert et al., 1997). Alternatively, we found two adjacent HBSs in the transferrin enhancer (Rolfs et al., 1997), and two or three adjacent HBSs were also found in the genes for phosphoglycerate kinase 1, enolase 1, lactate dehydrogenase A, glucose transporter-1, haem oxygenase 1, p35srj and insulin-like growth factor binding protein-1 (IGFBP-1) (for references, see Table 1).

Table 1.

Identified HIF-1 target genes

Identified HIF-1 target genes
Identified HIF-1 target genes

Several negative feedback regulatory pathways apparently exist to limit the hypoxic response. At the level of mRNA, we found a downregulation of HIF-1α mRNA concentrations by an unknown mechanism following prolonged exposure to hypoxia (Wenger et al., 1998c). At the transactivation level, HIF-1-dependent induction of p35srj has been reported, which in turn binds to the critical coactivator CBP/p300 and blocks the interaction with HIF-1α, resulting in a marked inhibition of transactivation (Bhattacharya et al., 1999).

HIF-1 target genes

Table 1 shows a compilation of the HIF-1 target genes identified so far. Only genes where functional HIF-1 DNA-binding or transactivation has been demonstrated are listed. These genes are involved in oxygen homeostasis at the cellular, local and systemic levels. Erythropoietin activates erythropoiesis to enhance the systemic oxygen transport capacity. Because iron is a limiting factor in haem formation, erythropoiesis is sustained by increased expression of transferrin and transferrin receptor to enhance iron supply to erythroid cells.

At the local level, HIF-1 activates vascular endothelial growth factor, as well as one of its receptors (Flt-1), which induce angiogenesis leading to an increase in the vascular density and hence a decrease in the diffusion distance for oxygen. Local blood circulation is also controlled by modulation of the vascular tone through the production of NO (nitric oxide synthase), CO (haem oxygenase 1), endothelin-1, adrenomedullin or activation of the α1B-adrenergic receptor. It is notable that HIF-1 binding to the HREs of the endothelium-specific genes Flt-1 and endothelin-1 has not been directly demonstrated, and it seems possible that the relatively endothelium-specific HIF-2 might regulate these genes.

At the cellular level, loss of ATP production in mitochondria is compensated by anaerobic glycolysis. Therefore, glucose uptake (glucose transporters) and glycolysis (glycolytic enzymes, see Table 1) are upregulated by HIF-1. Besides hypoxia, both insulin and insulin-like growth factors (IGF)-1 and 2 induce HIF-1α expression (Agani and Semenza, 1998; Feldser et al., 1999; Zelzer et al., 1998). In addition to glucose transporters and glycolytic enzymes, HIF-1 activates IGF-2 and the IGF-binding proteins (IGFBPs) 1, 2 and 3 but not 4, 5 and 6 (Feldser et al., 1999; Tazuke et al., 1998). Other pleiotropic growth factors and cytokines capable of inducing HIF-1 include epidermal growth factor, fibroblast growth factor-2, interleukin-1β and tumour necrosis factor-α (Feldser et al., 1999; Hellwig-Burgel et al., 1999). Thus, these results indicate a complex interplay between positive and negative regulatory growth factors, although further experiments are required to provide a more detailed picture.

Fig. 1 summarizes schematically the current models and hypotheses about how a cell senses, signals and reacts to reductions in oxygen concentration. The identification of the oxygen sensor and the corresponding signal-transduction pathway that leads to the activation of HIFs will be of principal research interest for the near future. With regard to the importance of hypoxia in many major diseases (cardiovascular defects, tumour growth), the findings will be of major importance for the treatment of these diseases.

Fig. 1.

Hypothetical model of oxygen sensing, signalling and gene regulation. Under hypoxic conditions, the ODD/TA domain of HIF-1α (blue) is stable and allows nuclear translocation and gene regulation. Under normoxic conditions, the ODD/TA domain (orange) obtains a signal from an unknown oxygen sensor which targets HIF-1α for proteolytic degradation via ubiquitination by an incompletely characterized complex containing the pVHL tumour suppressor protein. See text for details. ARNT, aryl hydrocarbon receptor (AhR) nuclear translocator; bHLH, basic-helix-loop-helix; CBP, CREB binding protein (CBP and p300 are two related transcriptional coactivators with histone acetyltransferase activity); Cul2, cullin 2; E2, unknown ubiquitin-conjugating enzyme; HBS, HIF-1 binding site; HRE, hypoxia-response element; ODD, oxygen-dependent degradation domain; PAS, Per-AhR/ARNT-Sim domain (see text); pol II, RNA polymerase II; pVHL, von-Hippel-Lindau tumour suppressor protein; Rbx1, a RING-H2 finger protein; TA, transactivation domain; Ub, ubiquitin; +1 indicates the transcriptional start site.

Fig. 1.

Hypothetical model of oxygen sensing, signalling and gene regulation. Under hypoxic conditions, the ODD/TA domain of HIF-1α (blue) is stable and allows nuclear translocation and gene regulation. Under normoxic conditions, the ODD/TA domain (orange) obtains a signal from an unknown oxygen sensor which targets HIF-1α for proteolytic degradation via ubiquitination by an incompletely characterized complex containing the pVHL tumour suppressor protein. See text for details. ARNT, aryl hydrocarbon receptor (AhR) nuclear translocator; bHLH, basic-helix-loop-helix; CBP, CREB binding protein (CBP and p300 are two related transcriptional coactivators with histone acetyltransferase activity); Cul2, cullin 2; E2, unknown ubiquitin-conjugating enzyme; HBS, HIF-1 binding site; HRE, hypoxia-response element; ODD, oxygen-dependent degradation domain; PAS, Per-AhR/ARNT-Sim domain (see text); pol II, RNA polymerase II; pVHL, von-Hippel-Lindau tumour suppressor protein; Rbx1, a RING-H2 finger protein; TA, transactivation domain; Ub, ubiquitin; +1 indicates the transcriptional start site.

I wish to thank C. Bauer, G. Camenisch, M. Gassmann and A. Scheid for helpful discussions and critical reading of the manuscript, C. Gasser for the artwork and the Sondermassnahmen des Bundes zur Förderung des akademischen Nachwuchses and the National Science Foundation (Grant 3156743.99) for financial support.

Acker
,
H.
(
1994
).
Mechanisms and meaning of cellular oxygen sensing in the organism
.
Respir. Physiol
.
95
,
1
10
.
Acker
,
H.
,
Dufau
,
E.
,
Huber
,
J.
and
Sylvester
,
D.
(
1989
).
Indications to an NADPH oxidase as a possible sensor in the rat carotid body
.
FEBS Lett
.
256
,
75
78
.
Agani
,
F.
and
Semenza
,
G. L.
(
1998
).
Mersalyl is a novel inducer of vascular endothelial growth factor gene expression and hypoxia-inducible factor 1 activity
.
Mol. Pharmac
.
54
,
749
754
.
Arany
,
Z.
,
Huang
,
L. E.
,
Eckner
,
R.
,
Bhattacharya
,
S.
,
Jiang
,
C.
,
Goldberg
,
M. A.
,
Bunn
,
H. F.
and
Livingston
,
D. M.
(
1996
).
An essential role for p300/CBP in the cellular response to hypoxia
.
Proc. Natl. Acad. Sci. USA
93
,
12969
12973
.
Archer
,
S. L.
,
Reeve
,
H. L.
,
Michelakis
,
E.
,
Puttagunta
,
L.
,
Waite
,
R.
,
Nelson
,
D. P.
,
Dinauer
,
M. C.
and
Weir
,
E. K.
(
1999
).
O2 sensing is preserved in mice lacking the gp91phox subunit of NADPH oxidase
.
Proc. Natl. Acad. Sci. USA
96
,
7944
7949
.
Bhattacharya
,
S.
,
Michels
,
C. L.
,
Leung
,
M. K.
,
Arany
,
Z. P.
,
Kung
,
A. L.
and
Livingston
,
D. M.
(
1999
).
Functional role of p35srj, a novel p300/CBP binding protein, during transactivation by HIF-1
.
Genes Dev
.
13
,
64
75
.
Budinger
,
G. R.
,
Duranteau
,
J.
,
Chandel
,
N. S.
and
Schumacker
,
P. T.
(
1998
).
Hibernation during hypoxia in cardiomyocytes. Role of mitochondria as the O2 sensor
.
J. Biol. Chem
.
273
,
3320
3326
.
Bunn
,
H. F.
and
Poyton
,
R. O.
(
1996
).
Oxygen sensing and molecular adaptation to hypoxia
.
Physiol. Rev
.
76
,
839
885
.
Canbolat
,
O.
,
Fandrey
,
J.
and
Jelkmann
,
W.
(
1998
).
Effects of modulators of the production and degradation of hydrogen peroxide on erythropoietin synthesis
.
Respir. Physiol
.
114
,
175
183
.
Carrero
,
P.
,
Okamoto
,
K.
,
Coumailleau
,
P.
,
O’brien
,
S.
,
Tanaka
,
H.
and
Poellinger
,
L.
(
2000
).
Redox-regulated recruitment of the transcriptional coactivators CREB-binding protein and SRC-1 to hypoxia-inducible factor 1α
.
Mol. Cell. Biol
.
20
,
402
415
.
Chandel
,
N. S.
,
Budinger
,
G. R.
,
Choe
,
S. H.
and
Schumacker
,
P. T.
(
1997
).
Cellular respiration during hypoxia. Role of cytochrome oxidase as the oxygen sensor in hepatocytes
.
J. Biol. Chem
.
272
,
18808
18816
.
Chandel
,
N. S.
,
Maltepe
,
E.
,
Goldwasser
,
E.
,
Mathieu
,
C. E.
,
Simon
,
M. C.
and
Schumacker
,
P. T.
(
1998
).
Mitochondrial reactive oxygen species trigger hypoxia-induced transcription
.
Proc. Natl. Acad. Sci. USA
95
,
11715
11720
.
Chilov
,
D.
,
Camenisch
,
G.
,
Kvietikova
,
I.
,
Ziegler
,
U.
,
Gassmann
,
M.
and
Wenger
,
R. H.
(
1999
).
Induction and nuclear translocation of hypoxia-inducible factor-1 (HIF-1): heterodimerization with ARNT is not necessary for nuclear accumulation of HIF-1α
.
J. Cell Sci
.
112
,
1203
1212
.
Conrad
,
P. W.
,
Freeman
,
T. L.
,
Beitner-Johnson
,
D.
and
Millhorn
,
D. E.
(
1999
).
EPAS1 trans-activation during hypoxia requires p42/p44 MAPK
.
J. Biol. Chem
.
274
,
33709
33713
.
Cross
,
A. R.
,
Henderson
,
L.
,
Jones
,
O. T.
,
Delpiano
,
M. A.
,
Hentschel
,
J.
and
Acker
,
H.
(
1990
).
Involvement of an NAD(P)H oxidase as a sensor protein in the rat carotid body
.
Biochem. J
.
272
,
743
747
.
Damert
,
A.
,
Ikeda
,
E.
and
Risau
,
W.
(
1997
).
Activator-protein-1 binding potentiates the hypoxia-inducible factor-1-mediated hypoxia-induced transcriptional activation of vascular-endothelial growth factor expression in C6 glioma cells
.
Biochem. J
.
327
,
419
423
.
Duranteau
,
J.
,
Chandel
,
N. S.
,
Kulisz
,
A.
,
Shao
,
Z.
and
Schumacker
,
P. T.
(
1998
).
Intracellular signaling by reactive oxygen species during hypoxia in cardiomyocytes
.
J. Biol. Chem
.
273
,
11619
11624
.
Ebert
,
B. L.
and
Bunn
,
H. F.
(
1998
).
Regulation of transcription by hypoxia requires a multiprotein complex that includes hypoxia-inducible factor 1, an adjacent transcription factor and p300/CREB binding protein
.
Mol. Cell. Biol
.
18
,
4089
4096
.
Ebert
,
B. L.
,
Firth
,
J. D.
and
Ratcliffe
,
P. J.
(
1995
).
Hypoxia and mitochondrial inhibitors regulate expression of glucose transporter-1 via distinct cis-acting sequences
.
J. Biol. Chem
.
270
,
29083
29089
.
Eckhart
,
A. D.
,
Yang
,
N.
,
Xin
,
X.
and
Faber
,
J. E.
(
1997
).
Characterization of the α1B-adrenergic receptor gene promoter region and hypoxia regulatory elements in vascular smooth muscle
.
Proc. Natl. Acad. Sci. USA
94
,
9487
9492
.
Ehleben
,
W.
,
Porwol
,
T.
,
Fandrey
,
J.
and
Acker
,
H.
(
1998
).
The influence of phenobarbital on cytochromes and reactive oxygen species in erythropoietin producing HepG2 cells
.
FEBS Lett
.
440
,
343
347
.
Ema
,
M.
,
Hirota
,
K.
,
Mimura
,
J.
,
Abe
,
H.
,
Yodoi
,
J.
,
Sogawa
,
K.
,
Poellinger
,
L.
and
Fujii-Kuriyama
,
Y.
(
1999
).
Molecular mechanisms of transcription activation by HLF and HIF1-α in response to hypoxia: their stabilization and redox signal-induced interaction with CBP/p300
.
EMBO J
.
18
,
1905
1914
.
Ema
,
M.
,
Taya
,
S.
,
Yokotani
,
N.
,
Sogawa
,
K.
,
Matsuda
,
Y.
and
Fujii-Kuriyama
,
Y.
(
1997
).
A novel bHLH-PAS factor with close sequence similarity to hypoxia-inducible factor 1α regulates the VEGF expression and is potentially involved in lung and vascular development
.
Proc. Natl. Acad. Sci. USA
94
,
4273
4278
.
Estes
,
S. D.
,
Stoler
,
D. L.
and
Anderson
,
G. R.
(
1995
).
Anoxic induction of a sarcoma virus-related VL30 retrotransposon is mediated by a cis-acting element which binds hypoxia-inducible factor 1 and an anoxia-inducible factor
.
J. Virol
.
69
,
6335
6341
.
Fandrey
,
J.
,
Frede
,
S.
,
Ehleben
,
W.
,
Porwol
,
T.
,
Acker
,
H.
and
Jelkmann
,
W.
(
1997
).
Cobalt chloride and desferrioxamine antagonize the inhibition of erythropoietin production by reactive oxygen species
.
Kidney Int
.
51
,
492
496
.
Fandrey
,
J.
,
Frede
,
S.
and
Jelkmann
,
W.
(
1994
).
Role of hydrogen peroxide in hypoxia-induced erythropoietin production
.
Biochem. J
.
303
,
507
510
.
Fandrey
,
J.
,
Seydel
,
F. P.
,
Siegers
,
C. P.
and
Jelkmann
,
W.
(
1990
).
Role of cytochrome P450 in the control of the production of erythropoietin
.
Life Sci
.
47
,
127
134
.
Feldser
,
D.
,
Agani
,
F.
,
Iyer
,
N. V.
,
Pak
,
B.
,
Ferreira
,
G.
and
Semenza
,
G. L.
(
1999
).
Reciprocal positive regulation of hypoxia-inducible factor 1α and insulin-like growth factor 2
.
Cancer Res
.
59
,
3915
3918
.
Firth
,
J. D.
,
Ebert
,
B. L.
,
Pugh
,
C. W.
and
Ratcliffe
,
P. J.
(
1994
).
Oxygen-regulated control elements in the phosphoglycerate kinase 1 and lactate dehydrogenase A genes: similarities with the erythropoietin 3′ enhancer
.
Proc. Natl. Acad. Sci. USA
91
,
6496
6500
.
Firth
,
J. D.
,
Ebert
,
B. L.
and
Ratcliffe
,
P. J.
(
1995
).
Hypoxic regulation of lactate dehydrogenase A. Interaction between hypoxia-inducible factor 1 and cyclic AMP response elements
.
J. Biol. Chem
.
270
,
21021
21027
.
Flamme
,
I.
,
Frohlich
,
T.
,
von Reutern
,
M.
,
Kappel
,
A.
,
Damert
,
A.
and
Risau
,
W.
(
1997
).
HRF, a putative basic helix–loop–helix-PAS-domain transcription factor is closely related to hypoxia-inducible factor-1α and developmentally expressed in blood vessels
.
Mech. Dev
.
63
,
51
60
.
Forsythe
,
J. A.
,
Jiang
,
B. H.
,
Iyer
,
N. V.
,
Agani
,
F.
,
Leung
,
S. W.
,
Koos
,
R. D.
and
Semenza
,
G. L.
(
1996
).
Activation of vascular endothelial growth factor gene transcription by hypoxia-inducible factor 1
.
Mol. Cell. Biol
.
16
,
4604
4613
.
Galson
,
D. L.
,
Tsuchiya
,
T.
,
Tendler
,
D. S.
,
Huang
,
L. E.
,
Ren
,
Y.
,
Ogura
,
T.
and
Bunn
,
H. F.
(
1995
).
The orphan receptor hepatic nuclear factor 4 functions as a transcriptional activator for tissue-specific and hypoxia-specific erythropoietin gene expression and is antagonized by EAR3/COUP-TF1
.
Mol. Cell. Biol
.
15
,
2135
2144
.
Gassmann
,
M.
,
Kvietikova
,
I.
,
Rolfs
,
A.
and
Wenger
,
R. H.
(
1997
).
Oxygen- and dioxin-regulated gene expression in mouse hepatoma cells
.
Kidney Int
.
51
,
567
574
.
Gerber
,
H. P.
,
Condorelli
,
F.
,
Park
,
J.
and
Ferrara
,
N.
(
1997
).
Differential transcriptional regulation of the two vascular endothelial growth factor receptor genes. Flt-1, but not Flk-1/KDR, is up-regulated by hypoxia
.
J. Biol. Chem
.
272
,
23659
23667
.
Gleadle
,
J. M.
,
Ebert
,
B. L.
and
Ratcliffe
,
P. J.
(
1995
).
Diphenylene iodonium inhibits the induction of erythropoietin and other mammalian genes by hypoxia. Implications for the mechanism of oxygen sensing
.
Eur. J. Biochem
.
234
,
92
99
.
Goldberg
,
M. A.
,
Dunning
,
S. P.
and
Bunn
,
H. F.
(
1988
).
Regulation of the erythropoietin gene: evidence that the oxygen sensor is a heme protein
.
Science
242
,
1412
1415
.
Goldwasser
,
E.
,
Alibali
,
P.
and
Gardner
,
A.
(
1995
).
Differential inhibition by iodonium compounds of induced erythropoietin expression
.
J. Biol. Chem
.
270
,
2628
2629
.
Gong
,
W.
,
Hao
,
B.
,
Mansy
,
S. S.
,
Gonzalez
,
G.
,
Gilles-Gonzalez
,
M. A.
and
Chan
,
M. K.
(
1998
).
Structure of a biological oxygen sensor: a new mechanism for heme-driven signal transduction
.
Proc. Natl. Acad. Sci. USA
95
,
15177
15182
.
Görlach
,
A.
,
Fandrey
,
J.
,
Holtermann
,
G.
and
Acker
,
H.
(
1994
).
Effects of cobalt on haem proteins of erythropoietin-producing HepG2 cells in multicellular spheroid culture
.
FEBS Lett
.
348
,
216
218
.
Görlach
,
A.
,
Holtermann
,
G.
,
Jelkmann
,
W.
,
Hancock
,
J. T.
,
Jones
,
S. A.
,
Jones
,
O. T.
and
Acker
,
H.
(
1993
).
Photometric characteristics of haem proteins in erythropoietin-producing hepatoma cells (HepG2)
.
Biochem. J
.
290
,
771
776
.
Gradin
,
K.
,
McGuire
,
J.
,
Wenger
,
R. H.
,
Kvietikova
,
I.
,
Whitelaw
,
M. L.
,
Toftgård
,
R.
,
Tora
,
L.
,
Gassmann
,
M.
and
Poellinger
,
L.
(
1996
).
Functional interference between hypoxia and dioxin signal transduction pathways: competition for recruitment of the Arnt transcription factor
.
Mol. Cell. Biol
.
16
,
5221
5231
.
Graven
,
K. K.
,
Yu
,
Q.
,
Pan
,
D.
,
Roncarati
,
J. S.
and
Farber
,
H. W.
(
1999
).
Identification of an oxygen responsive enhancer element in the glyceraldehyde-3-phosphate dehydrogenase gene
.
Biochim. Biophys. Acta
1447
,
208
218
.
Gu
,
Y. Z.
,
Moran
,
S. M.
,
Hogenesch
,
J. B.
,
Wartman
,
L.
and
Bradfield
,
C. A.
(
1998
).
Molecular characterization and chromosomal localization of a third α-class hypoxia inducible factor subunit, HIF3α.
Gene Expr
.
7
,
205
213
.
Harder
,
D. R.
,
Narayanan
,
J.
,
Birks
,
E. K.
,
Liard
,
J. F.
,
Imig
,
J. D.
,
Lombard
,
J. H.
,
Lange
,
A. R.
and
Roman
,
R. J.
(
1996
).
Identification of a putative microvascular oxygen sensor
.
Circ. Res
.
79
,
54
61
.
Hellwig-Burgel
,
T.
,
Rutkowski
,
K.
,
Metzen
,
E.
,
Fandrey
,
J.
and
Jelkmann
,
W.
(
1999
).
Interleukin-1β and tumor necrosis factor-α stimulate DNA binding of hypoxia-inducible factor-1
.
Blood
94
,
1561
1567
.
Ho
,
V. T.
and
Bunn
,
H. F.
(
1996
).
Effects of transition metals on the expression of the erythropoietin gene: further evidence that the oxygen sensor is a heme protein
.
Biochem. Biophys. Res. Commun
.
223
,
175
180
.
Hoffman
,
E. C.
,
Reyes
,
H.
,
Chu
,
F. F.
,
Sander
,
F.
,
Conley
,
L. H.
,
Brooks
,
B. A.
and
Hankinson
,
O.
(
1991
).
Cloning of a factor required for activity of the Ah (dioxin) receptor
.
Science
252
,
954
958
.
Hogenesch
,
J. B.
,
Chan
,
W. K.
,
Jackiw
,
V. H.
,
Brown
,
R. C.
,
Gu
,
Y. Z.
,
Pray-Grant
,
M.
,
Perdew
,
G. H.
and
Bradfield
,
C. A.
(
1997
).
Characterization of a subset of the basic-helix–loop–helix-PAS superfamily that interacts with components of the dioxin signaling pathway
.
J. Biol. Chem
.
272
,
8581
8593
.
Hogenesch
,
J. B.
,
Gu
,
Y. Z.
,
Jain
,
S. J.
and
Bradfield
,
C. A.
(
1998
).
The basic-helix–loop–helix-PAS orphan MOP3 forms transcriptionally active complexes with circadian and hypoxia factors
.
Proc. Natl. Acad. Sci. USA
95
,
5474
5479
.
Hohler
,
B.
,
Lange
,
B.
,
Holzapfel
,
B.
,
Goldenberg
,
A.
,
Hanze
,
J.
,
Sell
,
A.
,
Testan
,
H.
,
Moller
,
W.
and
Kummer
,
W.
(
1999
).
Hypoxic upregulation of tyrosine hydroxylase gene expression is paralleled, but not induced, by increased generation of reactive oxygen species in PC12 cells
.
FEBS Lett
.
457
,
53
56
.
Hu
,
B.
,
Wright
,
E.
,
Campbell
,
L.
and
Blanchard
,
K. L.
(
1997
).
In vivo analysis of DNA–protein interactions on the human erythropoietin enhancer
.
Mol. Cell. Biol
.
17
,
851
856
.
Hu
,
J.
,
Discher
,
D. J.
,
Bishopric
,
N. H.
and
Webster
,
K. A.
(
1998
).
Hypoxia regulates expression of the endothelin-1 gene through a proximal hypoxia-inducible factor-1 binding site on the antisense strand
.
Biochem. Biophys. Res. Commun
.
245
,
894
899
.
Huang
,
L. E.
,
Gu
,
J.
,
Schau
,
M.
and
Bunn
,
H. F.
(
1998
).
Regulation of hypoxia-inducible factor 1α is mediated by an O2-dependent degradation domain via the ubiquitin–proteasome pathway
.
Proc. Natl. Acad. Sci. USA
95
,
7987
7992
.
Huang
,
L. E.
,
Arany
,
Z.
,
Livingston
,
D. M.
and
Bunn
,
H. F.
(
1996
).
Activation of hypoxia-inducible transcription factor depends primarily upon redox-sensitive stabilization of its α subunit
.
J. Biol. Chem
.
271
,
32253
32259
.
Iwai
,
K.
,
Drake
,
S. K.
,
Wehr
,
N. B.
,
Weissman
,
A. M.
,
LaVaute
,
T.
,
Minato
,
N.
,
Klausner
,
R. D.
,
Levine
,
R. L.
and
Rouault
,
T. A.
(
1998
).
Iron-dependent oxidation, ubiquitination and degradation of iron regulatory protein 2: implications for degradation of oxidized proteins
.
Proc. Natl. Acad. Sci. USA
95
,
4924
4928
.
Iyer
,
N. V.
,
Kotch
,
L. E.
,
Agani
,
F.
,
Leung
,
S. W.
,
Laughner
,
E.
,
Wenger
,
R. H.
,
Gassmann
,
M.
,
Gearhart
,
J. D.
,
Lawler
,
A. M.
,
Yu
,
A. Y.
and
Semenza
,
G. L.
(
1998
).
Cellular and developmental control of O2 homeostasis by hypoxia-inducible factor 1α.
Genes Dev
.
12
,
149
162
.
Jiang
,
B. H.
,
Semenza
,
G. L.
,
Bauer
,
C.
and
Marti
,
H. H.
(
1996
).
Hypoxia-inducible factor 1 levels vary exponentially over a physiologically relevant range of O2 tension
.
Am. J. Physiol
.
271
,
C1172
C1180
.
Jiang
,
B. H.
,
Zheng
,
J. Z.
,
Leung
,
S. W.
,
Roe
,
R.
and
Semenza
,
G. L.
(
1997
).
Transactivation and inhibitory domains of hypoxia-inducible factor 1α. Modulation of transcriptional activity by oxygen tension
.
J. Biol. Chem
.
272
,
19253
19260
.
Kallio
,
P. J.
,
Okamoto
,
K.
,
O’brien
,
S.
,
Carrero
,
P.
,
Makino
,
Y.
,
Tanaka
,
H.
and
Poellinger
,
L.
(
1998
).
Signal transduction in hypoxic cells: inducible nuclear translocation and recruitment of the CBP/p300 coactivator by the hypoxia-inducible factor-1α.
EMBO J
.
17
,
6573
6586
.
Kallio
,
P. J.
,
Pongratz
,
I.
,
Gradin
,
K.
,
McGuire
,
J.
and
Poellinger
,
L.
(
1997
).
Activation of hypoxia-inducible factor 1α: posttranscriptional regulation and conformational change by recruitment of the Arnt transcription factor
.
Proc. Natl. Acad. Sci. USA
94
,
5667
5672
.
Kallio
,
P. J.
,
Wilson
,
W. J. S. O. B.
,
Makino
,
Y.
and
Poellinger
,
L.
(
1999
).
Regulation of the hypoxia-inducible transcription factor 1α by the ubiquitin-proteasome pathway
.
J. Biol. Chem
.
274
,
6519
6525
.
Kietzmann
,
T.
,
Freimann
,
S.
,
Bratke
,
J.
and
Jungermann
,
K.
(
1996
).
Regulation of the gluconeogenic phosphoenolpyruvate carboxykinase and glycolytic aldolase A gene expression by O2 in rat hepatocyte cultures. Involvement of hydrogen peroxide as mediator in the response to O2
.
FEBS Lett
.
388
,
228
232
.
Kietzmann
,
T.
,
Porwol
,
T.
,
Zierold
,
K.
,
Jungermann
,
K.
and
Acker
,
H.
(
1998
).
Involvement of a local Fenton reaction in the reciprocal modulation by O2 of the glucagon-dependent activation of the phosphoenolpyruvate carboxykinase gene and the insulin-dependent activation of the glucokinase gene in rat hepatocytes
.
Biochem. J
.
335
,
425
432
.
Kietzmann
,
T.
,
Roth
,
U.
,
Freimann
,
S.
and
Jungermann
,
K.
(
1997
).
Arterial oxygen partial pressures reduce the insulin-dependent induction of the perivenously located glucokinase in rat hepatocyte cultures: mimicry of arterial oxygen pressures by H2O2
.
Biochem. J
.
321
,
17
20
.
Kietzmann
,
T.
,
Roth
,
U.
and
Jungermann
,
K.
(
1999
).
Induction of the plasminogen activator inhibitor-1 gene expression by mild hypoxia via a hypoxia response element binding the hypoxia-inducible factor-1 in rat hepatocytes
.
Blood
94
,
4177
4185
.
Kotch
,
L. E.
,
Iyer
,
N. V.
,
Laughner
,
E.
and
Semenza
,
G. L.
(
1999
).
Defective vascularization of HIF-1α-null embryos is not associated with VEGF deficiency but with mesenchymal cell death
.
Dev. Biol
.
209
,
254
267
.
Kroll
,
S. L.
and
Czyzyk-Krzeska
,
M. F.
(
1998
).
Role of H2O2 and heme-containing O2 sensors in hypoxic regulation of tyrosine hydroxylase gene expression
.
Am. J. Physiol
.
274
,
C167
C174
.
Kummer
,
W.
and
Acker
,
H.
(
1995
).
Immunohistochemical demonstration of four subunits of neutrophil NAD(P)H oxidase in type I cells of carotid body
.
J. Appl. Physiol
.
78
,
1904
1909
.
Kvietikova
,
I.
,
Wenger
,
R. H.
,
Marti
,
H. H.
and
Gassmann
,
M.
(
1995
).
The transcription factors ATF-1 and CREB-1 bind constitutively to the hypoxia-inducible factor-1 (HIF-1) DNA recognition site
.
Nucleic Acids Res
.
23
,
4542
4550
.
Kvietikova
,
I.
,
Wenger
,
R. H.
,
Marti
,
H. H.
and
Gassmann
,
M.
(
1997
).
The hypoxia-inducible factor-1 DNA recognition site is cAMP-responsive
.
Kidney Int
.
51
,
564
566
.
Lahiri
,
S.
,
Buerk
,
D. G.
,
Chugh
,
D.
,
Osanai
,
S.
and
Mokashi
,
A.
(
1995
).
Reciprocal photolabile O2 consumption and chemoreceptor excitation by carbon monoxide in the cat carotid body: evidence for cytochrome a3 as the primary O2 sensor
.
Brain Res
.
684
,
194
200
.
Lahiri
,
S.
,
Ehleben
,
W.
and
Acker
,
H.
(
1999
).
Chemoreceptor discharges and cytochrome redox changes of the rat carotid body: role of heme ligands
.
Proc. Natl. Acad. Sci. USA
96
,
9427
9432
.
Lee
,
P. J.
,
Jiang
,
B. H.
,
Chin
,
B. Y.
,
Iyer
,
N. V.
,
Alam
,
J.
,
Semenza
,
G. L.
and
Choi
,
A. M.
(
1997
).
Hypoxia-inducible factor-1 mediates transcriptional activation of the heme oxygenase-1 gene in response to hypoxia
.
J. Biol. Chem
.
272
,
5375
5381
.
Levy
,
A. P.
,
Levy
,
N. S.
,
Wegner
,
S.
and
Goldberg
,
M. A.
(
1995
).
Transcriptional regulation of the rat vascular endothelial growth factor gene by hypoxia
.
J. Biol. Chem
.
270
,
13333
13340
.
Li
,
H.
,
Ko
,
H. P.
and
Whitlock
,
J. P.
(
1996
).
Induction of phosphoglycerate kinase 1 gene expression by hypoxia. Roles of ARNT and HIF1α.
J. Biol. Chem
.
271
,
21262
21267
.
Liu
,
Y.
,
Cox
,
S. R.
,
Morita
,
T.
and
Kourembanas
,
S.
(
1995
).
Hypoxia regulates vascular endothelial growth factor gene expression in endothelial cells. Identification of a 5′ enhancer
.
Circ. Res
.
77
,
638
643
.
Lok
,
C. N.
and
Ponka
,
P.
(
1999
).
Identification of a hypoxia response element in the transferrin receptor gene
.
J. Biol. Chem
.
274
,
24147
24152
.
Maxwell
,
P. H.
,
Wiesener
,
M. S.
,
Chang
,
G. W.
,
Clifford
,
S. C.
,
Vaux
,
E. C.
,
Cockman
,
M. E.
,
Wykoff
,
C. C.
,
Pugh
,
C. W.
,
Maher
,
E. R.
and
Ratcliffe
,
P. J.
(
1999
).
The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygendependent proteolysis
.
Nature
399
,
271
275
.
Melillo
,
G.
,
Musso
,
T.
,
Sica
,
A.
,
Taylor
,
L. S.
,
Cox
,
G. W.
and
Varesio
,
L.
(
1995
).
A hypoxia-responsive element mediates a novel pathway of activation of the inducible nitric oxide synthase promoter
.
J. Exp. Med
.
182
,
1683
1693
.
Nguyen
,
S. V.
and
Claycomb
,
W. C.
(
1999
).
Hypoxia regulates the expression of the adrenomedullin and HIF-1 genes in cultured HL-1 cardiomyocytes
.
Biochem. Biophys. Res. Commun
.
265
,
382
386
.
Okino
,
S. T.
,
Chichester
,
C. H.
and
Whitlock
,
J. P.
, Jr
(
1998
).
Hypoxia-inducible mammalian gene expression analyzed in vivo at a TATA-driven promoter and at an initiator-driven promoter
.
J. Biol. Chem
.
273
,
23837
23843
.
O’rourke
,
J. F.
,
Tian
,
Y. M.
,
Ratcliffe
,
P. J.
and
Pugh
,
C. W.
(
1999
).
Oxygen-regulated and transactivating domains in endothelial PAS protein 1: comparison with hypoxia-inducible factor-1α.
J. Biol. Chem
.
274
,
2060
2071
.
Palmer
,
L. A.
,
Semenza
,
G. L.
,
Stoler
,
M. H.
and
Johns
,
R. A.
(
1998
).
Hypoxia induces type II NOS gene expression in pulmonary artery endothelial cells via HIF-1
.
Am. J. Physiol
.
274
,
L212
L219
.
Porwol
,
T.
,
Ehleben
,
W.
,
Zierold
,
K.
,
Fandrey
,
J.
and
Acker
,
H.
(
1998
).
The influence of nickel and cobalt on putative members of the oxygen-sensing pathway of erythropoietin-producing HepG2 cells
.
Eur. J. Biochem
.
256
,
16
23
.
Pugh
,
C. W.
,
O’rourke
,
J. F.
,
Nagao
,
M.
,
Gleadle
,
J. M.
and
Ratcliffe
,
P. J.
(
1997
).
Activation of hypoxia-inducible factor-1; definition of regulatory domains within the α subunit
.
J. Biol. Chem
.
272
,
11205
11214
.
Qi
,
Y.
,
Jamindar
,
T. M.
and
Dawson
,
G.
(
1995
).
Hypoxia alters iron homeostasis and induces ferritin synthesis in oligodendrocytes
.
J. Neurochem
.
64
,
2458
2464
.
Richard
,
D. E.
,
Berra
,
E.
,
Gothie
,
E.
,
Roux
,
D.
and
Pouyssegur
,
J.
(
1999
).
p42/p44 mitogen-activated protein kinases phosphorylate hypoxia-inducible factor 1α (HIF-1α) and enhance the transcriptional activity of HIF-1
.
J. Biol. Chem
.
274
,
32631
32637
.
Rolfs
,
A.
,
Kvietikova
,
I.
,
Gassmann
,
M.
and
Wenger
,
R. H.
(
1997
).
Oxygen-regulated transferrin expression is mediated by hypoxia-inducible factor-1
.
J. Biol. Chem
.
272
,
20055
20062
.
Ryan
,
H. E.
,
Lo
,
J.
and
Johnson
,
R. S.
(
1998
).
HIF-1α is required for solid tumor formation and embryonic vascularization
.
EMBO J
.
17
,
3005
3015
.
Salceda
,
S.
,
Beck
,
I.
and
Caro
,
J.
(
1996
).
Absolute requirement of aryl hydrocarbon receptor nuclear translocator protein for gene activation by hypoxia
.
Arch. Biochem. Biophys
.
334
,
389
394
.
Salceda
,
S.
,
Beck
,
I.
,
Srinivas
,
V.
and
Caro
,
J.
(
1997
).
Complex role of protein phosphorylation in gene activation by hypoxia
.
Kidney Int
.
51
,
556
559
.
Salceda
,
S.
and
Caro
,
J.
(
1997
).
Hypoxia-inducible factor 1α (HIF-1α) protein is rapidly degraded by the ubiquitin–proteasome system under normoxic conditions. Its stabilization by hypoxia depends on redox-induced changes
.
J. Biol. Chem
.
272
,
22642
22647
.
Schmaltz
,
C.
,
Hardenbergh
,
P. H.
,
Wells
,
A.
and
Fisher
,
D. E.
(
1998
).
Regulation of proliferation–survival decisions during tumor cell hypoxia
.
Mol. Cell. Biol
.
18
,
2845
2854
.
Semenza
,
G. L.
(
1998
).
Hypoxia-inducible factor 1: master regulator of O2 homeostasis
.
Curr. Opin. Genet. Dev
.
8
,
588
594
.
Semenza
,
G. L.
,
Jiang
,
B. H.
,
Leung
,
S. W.
,
Passantino
,
R.
,
Concordet
,
J. P.
,
Maire
,
P.
and
Giallongo
,
A.
(
1996
).
Hypoxia response elements in the aldolase A, enolase 1 and lactate dehydrogenase A gene promoters contain essential binding sites for hypoxia-inducible factor 1
.
J. Biol. Chem
.
271
,
32529
32537
.
Semenza
,
G. L.
,
Roth
,
P. H.
,
Fang
,
H. M.
and
Wang
,
G. L.
(
1994
).
Transcriptional regulation of genes encoding glycolytic enzymes by hypoxia-inducible factor 1
.
J. Biol. Chem
.
269
,
23757
23763
.
Srinivas
,
V.
,
Zhang
,
L. P.
,
Zhu
,
X. H.
and
Caro
,
J.
(
1999a
).
Characterization of an oxygen/redox-dependent degradation domain of hypoxia-inducible factor α (HIF-α) proteins
.
Biochem. Biophys. Res. Commun
.
260
,
557
561
.
Srinivas
,
V.
,
Zhu
,
X.
,
Salceda
,
S.
,
Nakamura
,
R.
and
Caro
,
J.
(
1999b
).
Hypoxia-inducible factor 1α (HIF-1α) is a non-heme iron protein. Implications for oxygen sensing [retraction of V. Srinivas, X. Zhu, S. Salceda, R. Nakamura and J. Caro (1998). J. Biol. Chem. 273, 18019–18022]
.
J. Biol. Chem
.
274
,
1180
.
Suzuki
,
Y. J.
,
Forman
,
H. J.
and
Sevanian
,
A.
(
1997
).
Oxidants as stimulators of signal transduction
.
Free Radical Biol. Med
.
22
,
269
285
.
Tacchini
,
L.
,
Bianchi
,
L.
,
Bernelli-Zazzera
,
A.
and
Cairo
,
G.
(
1999
).
Transferrin receptor induction by hypoxia. HIF-1-mediated transcriptional activation and cell-specific post-transcriptional regulation
.
J. Biol. Chem
.
274
,
24142
24146
.
Tan
,
C. C.
and
Ratcliffe
,
P. J.
(
1991
).
Effect of inhibitors of oxidative phosphorylation on erythropoietin mRNA in isolated perfused rat kidneys
.
Am. J. Physiol
.
261
,
F982
F987
.
Taylor
,
B. L.
and
Zhulin
,
I. B.
(
1999
).
PAS domains: internal sensors of oxygen, redox potential and light
.
Microbiol. Mol. Biol. Rev
.
63
,
479
506
.
Taylor
,
C. T.
,
Lisco
,
S. J.
,
Awtrey
,
C. S.
and
Colgan
,
S. P.
(
1998
).
Hypoxia inhibits cyclic nucleotide-stimulated epithelial ion transport: role for nucleotide cyclases as oxygen sensors
.
J. Pharmac. Exp. Ther
.
284
,
568
575
.
Tazuke
,
S. I.
,
Mazure
,
N. M.
,
Sugawara
,
J.
,
Carland
,
G.
,
Faessen
,
G. H.
,
Suen
,
L. F.
,
Irwin
,
J. C.
,
Powell
,
D. R.
,
Giaccia
,
A. J.
and
Giudice
,
L. C.
(
1998
).
Hypoxia stimulates insulin-like growth factor binding protein 1 (IGFBP-1) gene expression in HepG2 cells: a possible model for IGFBP-1 expression in fetal hypoxia
.
Proc. Natl. Acad. Sci. USA
95
,
10188
10193
.
Tian
,
H.
,
Hammer
,
R. E.
,
Matsumoto
,
A. M.
,
Russell
,
D. W.
and
McKnight
,
S. L.
(
1998
).
The hypoxia-responsive transcription factor EPAS1 is essential for catecholamine homeostasis and protection against heart failure during embryonic development
.
Genes Dev
.
12
,
3320
3324
.
Tian
,
H.
,
McKnight
,
S. L.
and
Russell
,
D. W.
(
1997
).
Endothelial PAS domain protein 1 (EPAS1), a transcription factor selectively expressed in endothelial cells
.
Genes Dev
.
11
,
72
82
.
Vanden Hoek
,
T. L.
,
Becker
,
L. B.
,
Shao
,
Z.
,
Li
,
C.
and
Schumacker
,
P. T.
(
1998
).
Reactive oxygen species released from mitochondria during brief hypoxia induce preconditioning in cardiomyocytes
.
J. Biol. Chem
.
273
,
18092
18098
.
Wang
,
D.
,
Youngson
,
C.
,
Wong
,
V.
,
Yeger
,
H.
,
Dinauer
,
M. C.
,
Vega-Saenz Miera
,
E.
,
Rudy
,
B.
and
Cutz
,
E.
(
1996
).
NADPH-oxidase and a hydrogen peroxide-sensitive K+ channel may function as an oxygen sensor complex in airway chemoreceptors and small cell lung carcinoma cell lines
.
Proc. Natl. Acad. Sci. USA
93
,
13182
13187
.
Wang
,
G. L.
,
Jiang
,
B. H.
,
Rue
,
E. A.
and
Semenza
,
G. L.
(
1995a
).
Hypoxia-inducible factor 1 is a basic-helix–loop–helix-PAS heterodimer regulated by cellular O2 tension
.
Proc. Natl. Acad. Sci. USA
92
,
5510
5514
.
Wang
,
G. L.
,
Jiang
,
B. H.
and
Semenza
,
G. L.
(
1995b
).
Effect of protein kinase and phosphatase inhibitors on expression of hypoxia-inducible factor 1
.
Biochem. Biophys. Res. Commun
.
216
,
669
675
.
Wang
,
G. L.
and
Semenza
,
G. L.
(
1993a
).
Characterization of hypoxia-inducible factor 1 and regulation of DNA binding activity by hypoxia
.
J. Biol. Chem
.
268
,
21513
21518
.
Wang
,
G. L.
and
Semenza
,
G. L.
(
1993b
).
Desferrioxamine induces erythropoietin gene expression and hypoxia-inducible factor 1 DNA-binding activity: implications for models of hypoxia signal transduction
.
Blood
82
,
3610
3615
.
Wenger
,
R. H.
and
Gassmann
,
M.
(
1997
).
Oxygen(es) and the hypoxia-inducible factor-1
.
Biol. Chem
.
378
,
609
616
.
Wenger
,
R. H.
,
Kvietikova
,
I.
,
Rolfs
,
A.
,
Camenisch
,
G.
and
Gassmann
,
M.
(
1998a
).
Oxygen-regulated erythropoietin gene expression is dependent on a CpG methylation-free hypoxia-inducible factor-1 DNA-binding site
.
Eur. J. Biochem
.
253
,
771
777
.
Wenger
,
R. H.
,
Marti
,
H. H.
,
Bauer
,
C.
and
Gassmann
,
M.
(
1998b
).
Optimal erythropoietin expression in human hepatoma cell lines requires activation of multiple signalling pathways
.
Int. J. Mol. Med
.
2
,
317
324
.
Wenger
,
R. H.
,
Marti
,
H. H.
,
Schuerer-Maly
,
C. C.
,
Kvietikova
,
I.
,
Bauer
,
C.
,
Gassmann
,
M.
and
Maly
,
F. E.
(
1996
).
Hypoxic induction of gene expression in chronic granulomatous disease-derived B-cell lines: oxygen sensing is independent of the cytochrome b558-containing nicotinamide adenine dinucleotide phosphate oxidase
.
Blood
87
,
756
761
.
Wenger
,
R. H.
,
Rolfs
,
A.
,
Kvietikova
,
I.
,
Spielmann
,
P.
,
Zimmermann
,
D. R.
and
Gassmann
,
M.
(
1997
).
The mouse gene for hypoxia-inducible factor-1α. Genomic organization, expression and characterization of an alternative first exon and 5′ flanking sequence
.
Eur. J. Biochem
.
246
,
155
165
.
Wenger
,
R. H.
,
Rolfs
,
A.
,
Spielmann
,
P.
,
Zimmermann
,
D. R.
and
Gassmann
,
M.
(
1998c
).
Mouse hypoxia-inducible factor-1α is encoded by two different mRNA isoforms – expression from a tissue-specific and a housekeeping-type promoter
.
Blood
91
,
3471
3480
.
Wiesener
,
M. S.
,
Turley
,
H.
,
Allen
,
W. E.
,
Willam
,
C.
,
Eckardt
,
K. U.
,
Talks
,
K. L.
,
Wood
,
S. M.
,
Gatter
,
K. C.
,
Harris
,
A. L.
,
Pugh
,
C. W.
,
Ratcliffe
,
P. J.
and
Maxwell
,
P. H.
(
1998
).
Induction of endothelial PAS domain protein-1 by hypoxia: Characterization and comparison with hypoxia-inducible factor-1α.
Blood
92
,
2260
2268
.
Wilson
,
D. F.
,
Mokashi
,
A.
,
Chugh
,
D.
,
Vinogradov
,
S.
,
Osanai
,
S.
and
Lahiri
,
S.
(
1994
).
The primary oxygen sensor of the cat carotid body is cytochrome a3 of the mitochondrial respiratory chain
.
FEBS Lett
.
351
,
370
374
.
Wood
,
S. M.
,
Gleadle
,
J. M.
,
Pugh
,
C. W.
,
Hankinson
,
O.
and
Ratcliffe
,
P. J.
(
1996
).
The role of the aryl hydrocarbon receptor nuclear translocator (ARNT) in hypoxic induction of gene expression. Studies in ARNT-deficient cells
.
J. Biol. Chem
.
271
,
15117
15123
.
Wood
,
S. M.
,
Wiesener
,
M. S.
,
Yeates
,
K. M.
,
Okada
,
N.
,
Pugh
,
C. W.
,
Maxwell
,
P. H.
and
Ratcliffe
,
P. J.
(
1998
).
Selection and analysis of a mutant cell line defective in the hypoxia-inducible factor-1 α-subunit (HIF-1α). Characterization of HIF-1α-dependent and -independent hypoxia-inducible gene expression
.
J. Biol. Chem
.
273
,
8360
8368
.
Youngson
,
C.
,
Nurse
,
C.
,
Yeger
,
H.
,
Curnutte
,
J. T.
,
Vollmer
,
C.
,
Wong
,
V.
and
Cutz
,
E.
(
1997
).
Immunocytochemical localization on O2-sensing protein (NADPH oxidase) in chemoreceptor cells
.
Microsc. Res. Tech
.
37
,
101
106
.
Youngson
,
C.
,
Nurse
,
C.
,
Yeger
,
H.
and
Cutz
,
E.
(
1993
).
Oxygen sensing in airway chemoreceptors
.
Nature
365
,
153
155
.
Yu
,
A. Y.
,
Shimoda
,
L. A.
,
Iyer
,
N. V.
,
Huso
,
D. L.
,
Sun
,
X.
,
McWilliams
,
R.
,
Beaty
,
T.
,
Sham
,
J. S.
,
Wiener
,
C. M.
,
Sylvester
,
J. T.
and
Semenza
,
G. L.
(
1999
).
Impaired physiological responses to chronic hypoxia in mice partially deficient for hypoxia-inducible factor 1α.
J. Clin. Invest
.
103
,
691
696
.
Zelzer
,
E.
,
Levy
,
Y.
,
Kahana
,
C.
,
Shilo
,
B. Z.
,
Rubinstein
,
M.
and
Cohen
,
B.
(
1998
).
Insulin induces transcription of target genes through the hypoxia-inducible factor HIF-1α/ARNT
.
EMBO J
.
17
,
5085
5094
.