IN HUMANS AND OTHER mammals, decreased oxygen tension triggers specific and tightly regulated cellular, vascular, and erythropoietic responses. An association between polycythemia and people living at high altitudes was first reported in 1863.1 Erythropoietin (Epo), a 34.4-kD glycoprotein hormone, was subsequently identified as the humoral regulator of red blood cell production. Decreased tissue oxygen tension modulates Epo levels by increasing expression of the Epo gene. Since the cloning of the Epo gene in 1985,2,3 considerable progress has been made in understanding the molecular mechanisms by which the Epo gene is regulated by environmental, tissue-specific, and developmental cues.

Erythropoiesis, which normally proceeds at a low basal level to replace aged red blood cells, is highly induced by loss of red blood cells, decreased ambient oxygen tension, increased oxygen affinity for hemoglobin, and other stimuli that decrease delivery of oxygen to the tissues. In states of severe hypoxia, production of Epo is increased up to 1,000-fold. The secreted hormone circulates in the blood and binds to receptors expressed specifically on erythroid progenitor cells, thereby promoting the viability, proliferation, and terminal differentiation of erythroid precursors, resulting in an increase in red blood cell mass. The oxygen carrying capacity of the blood is thus enhanced, increasing tissue oxygen tension, thereby completing the negative feedback loop (Fig 1).4,5 

Fig. 1.

Feedback loop responsible for the regulation of erythropoiesis.

Fig. 1.

Feedback loop responsible for the regulation of erythropoiesis.

Close modal

Research on the regulation of the Epo gene has shown a general system of oxygen-sensing, signaling, and transcriptional regulation of a broad range of physiologically relevant genes, including those encoding angiogenic growth factors, glucose transporters, and enzymes involved in adaptation to hypoxia.

Temporal and tissue-specific signals limit expression of the Epogene primarily to specific cells in the fetal liver and the adult kidney.

Localization of Epo production to the kidneys was first demonstrated by Jacobsen et al,6 who showed that, after bilateral nephrectomy, rats and rabbits do not respond to hemorrhage with an appropriate increase in plasma Epo levels. Erslev et al7have proposed that the peritubular region of the renal cortex is the ideal location for Epo production. Oxygen consumption in the kidney is determined largely by sodium reabsorption, which in turn depends on the filtered load. Because the glomerular filtration rate is roughly proportional to renal blood flow, renal oxygen consumption is linked to renal blood flow. Therefore, oxygen tension at the site of the Epo-producing cells is relatively independent of changes in renal blood flow. At high hematocrit levels, viscosity increases to the point that blood flow to tissues is compromised. If the primary site of Epo production were in an organ other than the kidney, the resultant decrease in tissue oxygen tension would lead to a vicious cycle of increasing erythropoiesis causing worsening hypoxia.

When kidney cells were separated into glomerular and tubular fractions,Epo mRNA was found only in the tubular fraction that included the peritubular interstitium.8 Consistent with these findings, in situ hybridization studies with 35S-labeledEpo RNA9,10 or DNA11 probes on kidney tissue from anemic mice showed Epo mRNA in peritubular interstitial cells. The number of these cells expressing EpomRNA increased with a decreasing hematocrit level.12 In contrast to these studies, other investigators have demonstrated Epo production by renal tubular cells using in situ hybridization forEpo mRNA,13 immunohistochemistry with Epo-specific antibodies,13 and detection of β-galactosidase in transgenic mice bearing a 7-kb fragment of the Epo gene linked to the lacZ gene.14 Human renal tumor cells of tubular origin can express Epo.15Immunohistochemistry using Epo-specific antibodies is confounded by the reabsorption of circulating Epo by renal tubular cells.

Two studies have demonstrated the colocalization of Epo-producing cells and immunoreactivity to 5′-ectonucleotidase, suggesting that the cells are likely to be fibroblasts.16,17 Maxwell et al17 prepared transgenic mice using regulatory sequences from the mouse Epo gene flanking the SV40 T antigen as a marker gene. In one line of these transgenic mice, the transgene was fortuitously integrated into the endogenous Epo locus by homologous recombination. This provided a model in which the marker gene, inserted into the Epo locus, is subject to all of the same tissue-specific controls as the endogenous Epo gene. Consequently, expression of the SV40 T antigen permitted immunohistochemical identification of the Epo-producing cells to be the fibroblast-like type I interstitial cells.17 When all of the somewhat contradictory studies cited above are carefully weighed, the bulk of convincing evidence favors a peritubular interstitial cell as the primary site of regulated Epo production in the kidney.

Before birth, Epo is primarily produced in the liver. The primary site of Epo production switches from liver to kidney shortly after birth,18,19 but the signals governing this change are poorly understood. In the liver, an oxygen gradient is established as oxygen-rich blood from the portal triads becomes depleted of oxygen as it flows towards the central vein. Consequently, in transgenic mice, both the Epo transgene and the endogenous Epo gene are preferentially expressed near the central vein where oxygen tension is lowest.10 

Two Epo-producing cell types were identified in the liver: hepatocytes and a nonparenchymal cell type.10,20 The identity of the nonparenchymal cells was established using transgenic mice bearing the SV40 T antigen homologously recombined into the Epo locus. In these mice, SV40 T antigen expression was observed in a subset of nonparenchymal cells, identified as Ito cells, as well as in a subset of hepatocytes.21 

Highly sensitive assays have shown low levels of Epo mRNA in the kidneys and livers of unstimulated mice and rats,9,22-24 consistent with low basal levels of Epo in serum. Low levels of Epo expression have also been detected in the lung, spleen, brain, and testis of rats.23,24 

Expression and production of both Epo and Epo-receptors has been demonstrated in the brain.23,25-27 Oxygen-regulated expression of Epo has been observed in astrocytes both in vitro in cultured astrocytes25,27 and in vivo.23,27 The presence of Epo receptors, the inability of Epo to cross the blood-brain barrier, and the regulated expression of Epo in the brain have led researchers to propose a paracrine function for Epo in neural tissue. Recent evidence demonstrates that Epo can protect neurons from ischemic damage in vivo.28 

The discovery that both Epo mRNA and Epo protein are expressed in erythroid progenitors29,30 has raised the intriguing possibility that tonic low-level erythropoiesis may be supported by autocrine stimulation, whereas circulating (hormonal) Epo provides a more robust stimulus to erythropoiesis during hypoxic stress.

For decades, a tissue culture model eluded investigators studying the regulation of Epo. Some cells, such as rat kidney mesangial cells,31 the renal cell line RC-1,32 and hepatic carcinomas,33 produced Epo at very low levels with minimal induction by hypoxia. Cell lines that produce significant amounts of Epo in a regulated fashion were discovered by screening a range of renal and hepatic cells in culture.34 Two human hepatoma cell lines, Hep3B and HepG2, were shown to produce significant amounts of Epo constitutively, with marked induction by hypoxia. The magnitude and time course of the induction of Epo mRNA paralleled Epo protein production. The discovery of a tissue culture model demonstrated that individual cells contain the apparatus necessary for oxygen sensing and the consequent regulation of gene expression. Hep3B and HepG2 cells have proved to be an invaluable tool in exploring the molecular basis of Epo gene regulation.

In addition to tissue-specific and developmental signals, Epogene expression is modulated by a number of physiological and pharmacological agents (Table 1). Regulation of Epo by hypoxia and other stimuli occurs at the mRNA level. In the kidneys of mice made anemic by blood loss,Epo mRNA was increased approximately 200-fold over the level in the kidneys of normal control mice.35,Epo mRNA was induced in the liver as well, but at a lower level of expression. The increase in Epo mRNA reached a maximum at 4 to 8 hours after induction. The magnitude of induction was proportional to the degree of anemia. Similarly, injection of cobalt chloride into rats inducedEpo expression in the kidney and, to a lesser degree, in the liver.36 The time course and level of induction of EpomRNA paralleled induction of Epo in serum measured by radioimmunoassay.36 

Table 1.

Agents That Affect Epo Gene Expression

Induction in vivo and in vitro  
 Hypoxia  
 Transition metals: Co2+, Ni2+, Mn2+ 118,192 
 Iron chelators: desferrioxamine120 
Abrogation of induction in Hep3B and HepG2 cells  
 Carbon monoxide91,118,193  
 Nitric oxide, nitroprusside91 
 Hydrogen peroxide86,194  
 Inflammatory cytokines: TNF-α, IL-1181,182  
 Phorbol esters: Forskolin193,195-197  
 Protein synthesis inhibitors: cycloheximide83,118  
Induction in vivo and in vitro  
 Hypoxia  
 Transition metals: Co2+, Ni2+, Mn2+ 118,192 
 Iron chelators: desferrioxamine120 
Abrogation of induction in Hep3B and HepG2 cells  
 Carbon monoxide91,118,193  
 Nitric oxide, nitroprusside91 
 Hydrogen peroxide86,194  
 Inflammatory cytokines: TNF-α, IL-1181,182  
 Phorbol esters: Forskolin193,195-197  
 Protein synthesis inhibitors: cycloheximide83,118  

This list is not comprehensive. Other agents have been shown either to activate or suppress HIF-1 and therefore would be expected to affect Epo gene expression.

Nuclear run-on assays using nuclei prepared from the kidneys of rats that had been made hypoxic or treated with cobalt showed an increase in transcription of the Epo gene.8 Similar results were obtained with Hep3B cells.37 Transcriptionally active nuclear extracts from hypoxic Hep3B cells were compared with extracts from cells cultured in normoxia.38 The hypoxic nuclear extracts supported a higher rate of Epo transcription in vitro, demonstrating the presence of hypoxically inducible trans-acting factors capable of interacting with cis-acting sequences from theEpo gene.

Comparison of the human and murine Epo genes provided clues to the location of key regulatory domains in the Epogene.39-41 Three noncoding segments of the Epogene are highly conserved between human and mouse sequences: the promoter, the first intron, and a 120-bp region 100 bp 3′ to the polyadenylation site.

Both transient transfection experiments and studies with transgenic mice have been used to identify functionally important cis-acting elements. Experiments with transgenic mice mapped broad regions of theEpo gene that regulate Epo expression in response to tissue-specific and developmental signals. Transient transfections of cultured cells have been used to characterize cis-acting sequences that are critical for the response to hypoxia. Conserved sequences both 5′ and 3′ of the Epo gene proved to be important for regulation of the Epo gene, but deletion of the conserved sequences in the first intron did not influence hypoxic induction.42 Regulatory elements in the Epo gene are portrayed in Fig 2.

Fig. 2.

Structure of the human Epo gene. Exons are indicated by solid black boxes; 5′ and 3′ untranslated regions are indicated by open rectangles. Areas of homology between human and murine noncoding sequences are shown with blue rectangles, and the region of liver specific DNase I hypersensitivity is shown with a green rectangle. The 3′ enhancer is expanded for greater detail. Sites that are functionally critical for hypoxic induction are underscored in red. Binding of HIF-1, HNF-4, and p300 is illustrated. As indicated by the arrow, p300 is capable of interacting with the basal transcriptional machinery in the promoter.

Fig. 2.

Structure of the human Epo gene. Exons are indicated by solid black boxes; 5′ and 3′ untranslated regions are indicated by open rectangles. Areas of homology between human and murine noncoding sequences are shown with blue rectangles, and the region of liver specific DNase I hypersensitivity is shown with a green rectangle. The 3′ enhancer is expanded for greater detail. Sites that are functionally critical for hypoxic induction are underscored in red. Binding of HIF-1, HNF-4, and p300 is illustrated. As indicated by the arrow, p300 is capable of interacting with the basal transcriptional machinery in the promoter.

Close modal
Transgenic experiments.

Transgenic mice were initially produced containing a 4-kb fragment that included the human Epo gene, 400 bp of 5′-flanking sequence, and 700 bp of 3′-flanking sequence. The transgene was widely expressed, causing the mice to become polycythemic, indicating that additional cis-acting sequences, required for tissue-specific and developmental regulation, lie outside this construct. Nevertheless, in the liver, the transgene was upregulated by anemia and cobalt chloride. Thus, cis-acting sequences capable of mediating oxygen-regulated gene expression in liver cells lie between 400 bp 5′ and 700 bp 3′ of the Epo gene.43 

Tissue-specific regulatory domains were mapped in subsequent transgenic experiments by use of constructs with varying lengths of Epoupstream flanking sequence. Promiscuous expression of the Epogene, seen in transgenic constructs with 300 bp44 or 400 bp43 of 5′ flanking sequence, was extinguished in constructs containing 6 kb of 5′ flanking sequence.45In constructs containing 9.5 kb or less of 5′ flanking sequence, inducible expression of the Epo transgene was observed in the liver but not in the kidney.44 However, physiological expression of the Epo transgene with inducible expression in the kidney was seen in transgenic mice containing 14 kb of 5′ flanking sequence.46 Thus, transgenic experiments indicate that a repressive element(s) exists between 0.4 and 6 kb upstream of the Epo gene, and a kidney-specific inducible element(s) exists between 9.5 and 14 kb upstream of the Epo gene.

Epo 3′ enhancer.

A liver-specific DNase I hypersensitivity site was discovered in the 3′ flanking sequence of an Epo transgene.47Analysis of this region of the Epo gene by transient transfections of reporter constructs led to the identification of a hypoxically inducible enhancer.47-50 In both the mouse and human Epo genes, this enhancer lies in a highly conserved region 120 bp 3′ to the polyadenylation site. As is typical of eukaryotic transcriptional enhancers, activity was independent of orientation and distance from the promoter. The enhancer demonstrated the same stimulus specificity as the Epo gene with responses to hypoxia, cobaltous chloride, and iron chelation, but not to cyanide and 2-deoxyglucose.

Detailed characterization of the Epo 3′ enhancer defined 3 sites that are critical for regulation by hypoxia.50-52On the 5′ side, the sequence CACGTGCT was the first response element to be characterized for the transcription factor, hypoxia inducible factor-1 (HIF-1).51 Binding of HIF-1 to this site is induced by hypoxia, and an intact HIF-1 binding site is necessary for hypoxically inducible function of the Epo enhancer. In addition to the hypoxically inducible DNA-binding activity, HIF-1, this site also binds another complex constitutively. The transcription factors ATF-1 and CREB-1 have been shown to be involved in this complex in vitro, but it is not clear whether these factors play a functional role in the Epo enhancer.53 Further details concerning the function and activation of HIF-1 are discussed below.

A second site, 7 bp 3′ to the HIF-1 site, has the sequence CACA in the human Epo gene. No proteins are known to bind to this site, but mutation of this site abrogates hypoxia inducible activity of the enhancer. A similar sequence has been found adjacent to HIF-1 sites in other genes.54 These first 2 sites (HIF-1 and CACA) require the presence of a third site for hypoxically inducible transcription, unless the enhancer is directly upstream from the promoter.52 

The sequence of the third site in the Epo enhancer is a direct repeat of 2 steroid hormone receptor half sites separated by 2 bp, termed a DR-2 site.50 Mutations of this site ablate or markedly inhibit hypoxic induction.50-52 Nuclear proteins from a broad range of cell types bind strongly to this site, as demonstrated by both electrophoretic mobility shift assays and DNase I footprinting experiments. However, binding of proteins to this site is not oxygen-dependent either in vivo or in vitro.50-52,55 In some non–Epo-producing cells, a complex does not form at this site in vivo.55 Hormones whose biological actions depend on binding to nuclear receptors had no effect on the hypoxic induction of a reporter gene containing the Epo promoter and Epoenhancer.50 These results suggested that the DR-2 site might bind an orphan nuclear receptor, a DNA-binding protein that shares structural homology with hormone binding nuclear receptors but lacks a known ligand. Screening a variety of in vitro-translated orphan receptors showed that HNF-4α bound specifically to this site.56 Hypoxic induction of Epo is abolished in Hep3B cells expressing a dominant negative mutant of HNF-4. HNF-4 is expressed in the renal cortex and liver, like Epo, as well as in the intestine. Thus, HNF-4 may contribute to the tissue specificity of Epo gene expression.

Promoter.

The Epo promoter does not have consensus TATA or CAAT elements in either the mouse or human genes. Comparison of the 5′ flanking sequences of the human and murine Epo genes shows 73% overall sequence identity and 8 areas of even higher homology.39,40 

The Epo promoter contributes to the hypoxic inducibility of theEpo gene.57 After deletion of the 3′ enhancer, expression of a stably transfected marked Epo gene was induced approximately 10-fold in response to hypoxia.42The minimal promoter acts synergistically with the 3′ enhancer to confer a 40-fold induction in response to hypoxia.50 

The minimal Epo promoter capable of induction by hypoxia encompasses 117 bp 5′ to the transcription initiation site.50 A segment of 17 bp (−61 to −45) is responsible for this upregulation by hypoxia.58 There is no HIF-1 consensus sequence at this site. Using computer homology matching, a HIF-1 site was identified at position −180 in theEpo 5′ flanking sequence.59 However, reporter gene experiments suggest that this site is not a functional hypoxia response element.50 GATA sites60,61 and a ribonucleoprotein binding site62 have also been described, but the role of these proteins in the response to hypoxia is not proven. Addition of antisense oligonucleotides to GATA elements increased Epo gene expression, whereas the addition of antisense oligonucleotides to CACCC elements decreased Epo gene expression, indicating that the Epo promoter is regulated negatively by GATA sites and positively by CACCC sites.60L-NMMA, a nitric oxide synthase inhibitor, was recently demonstrated to increase binding of GATA factors in the Epo promoter and decrease Epo expression.63 

Methylation of CpG sites in the Epo promoter varies between Epo-producing and non–Epo-producing cells. By inhibiting the formation of DNA-binding complexes and by the binding of inhibitory methyl-CpG binding proteins, methylation may contribute to tissue-specific activity of the Epo promoter.64 

Enhanced transcription accounts for most, but probably not all, of the hypoxic induction of the Epo gene. In Hep3B cells, nuclear run-on experiments showed about a 10-fold increase in transcription ofEpo mRNA during exposure to 1% O2 in the setting of a 50- to 100-fold increase in the steady-state level of EpomRNA.37 In 2 other hypoxically regulated genes, tyrosine hydroxylase (TH)65,66 and VEGF,67,68approximately 50% of the enhanced gene expression in response to hypoxia is due to increased mRNA stability. For both genes, specific mRNA binding proteins have been demonstrated in cytosolic extracts of hypoxic cells.66,68 

In comparison to TH and VEGF, less is known about posttranscriptional regulation of the Epo gene. Inhibitors of transcription markedly prolong the half-life of Epo mRNA, thus making actinomycin chase experiments uninterpretable.37 Two proteins, 70 and 135 kD, which have been designated EpomRNA-binding protein (ERBP), bind to a 120-bp pyrimidine-rich region in the 3′ UTR of Epo mRNA.69 This interaction does not seem to be regulated by oxygen tension,69 but binding is subject to redox control.70 Heat shock protein 70 participates in a complex with ERBP and Epo mRNA.71 Deletion of the ERBP binding site prolongs the half-life of Epo mRNA and eliminates hypoxically induced stabilization.72 Deletion of a 50-bp segment lying 70 bp downstream of the binding site for these proteins causes a 7-fold increase in the half-life of a transfected markedEpo gene.42 Epo, VEGF, and TH mRNA can cross-compete for binding in mobility shift assays, suggesting common features to the regulation of mRNA stability in these 3 genes.67,73 

The transcription factor HIF-1 mediates hypoxically inducible transcription of oxygen-regulated genes. The HIF-1 site in the Epo3′ enhancer is the primary element in the Epo gene that mediates the transcriptional response to hypoxia. The time course of HIF-1 activation mimics the induction of the Epogene.74 Hypoxia, cobalt, and DFO, stimuli that triggerEpo gene expression, also activate HIF-1 DNA binding. The HIF-1 site in the Epo 3′ enhancer was the first hypoxia response element (HRE) to be identified and was used for the affinity purification of HIF-1.75 HIF-1 is a heterodimer composed of 120-kD α and 91- to 94-kD β subunits, both of which are basic helix-loop-helix (bHLH) proteins in the PAS (Per-AHR-ARNT-SIM) family of transcription factors.76 HIF-1β is the previously cloned and characterized aryl hydrocarbon receptor nuclear translocator (ARNT), which forms a heterodimer with the aryl hydrocarbon receptor (AHR), mediating regulation of genes involved in the transcriptional response to xenobiotics and oxidant stress.77,78 

Both HIF-1α79 and ARNT80 have basic domains that are critical for DNA binding and PAS domains that are crucial for dimerization. The domain structure of HIF-1α is shown in Fig 3. Interaction between HIF-1α and ARNT requires both bHLH and PAS domains.81 Deletion of both the basic domain and the carboxy-terminal activation domains results in a dominant negative form of HIF-1α.79 Transactivation by the HIF-1 heterodimer requires the N-terminal DNA-binding and heterodimerization domains of ARNT, but not its C-terminal activation domain.82 

Fig. 3.

Structure of HIF-1. Open rectangles represent the PAS A and B domains. Solid black rectangles within the oxygen-dependent degradation domain indicate the location of PEST sequences.

Fig. 3.

Structure of HIF-1. Open rectangles represent the PAS A and B domains. Solid black rectangles within the oxygen-dependent degradation domain indicate the location of PEST sequences.

Close modal

HIF-1 activity has been demonstrated in a wide variety of cells by both electrophoretic mobility shift assays83 and transfections of a reporter gene containing HIF-1 sites.84 Steady-state levels of HIF-1α and ARNT mRNA are not significantly affected by oxygen tension.81,85,86 At the protein level, ARNT abundance is also not dependent on oxygen tension. In contrast, the HIF-1α subunit is only detectable in cells treated with hypoxia or stimuli that mimic hypoxia, such as cobalt or iron chelators. In normoxic cells or in cells pretreated with H2O2before deoxygenation, HIF-1α protein is barely detectable.86 Thus, HIF-1 activation correlates with the oxygen-dependent accumulation of HIF-1α protein. Under normoxic conditions, HIF-1α is rapidly degraded by the ubiquitin-proteasome pathway.87,88 Specific inhibitors of this pathway markedly increase HIF-1α abundance. Hypoxia and iron chelation dramatically increase the half-life of HIF-1α, permitting the formation of functionally active HIF-1α/ARNT heterodimers. A domain from the central region of HIF-1α was identified that is critical for oxygen-regulated protein degradation.88 This region spans amino acids 401-603 with further deletions causing a diminution of the magnitude of hypoxic inducibility.88 Sequences from within this region mediate hypoxically inducible transactivation.89,90 Deletion of this oxygen-dependent degradation (ODD) domain resulted in stabilization of HIF-1α and constitutive HIF-1 DNA-binding activity independent of oxygen tension.88 Furthermore, the ODD domain is transportable, ie, it is capable of conferring oxygen-dependent degradation on a heterologous protein.88 Both carbon monoxide and nitric oxide donors suppress degradation of HIF-1α via the ODD domain.91 

A second domain at the C-terminus of the HIF-1α protein (amino acids 775-826) was also shown to mediate hypoxically inducible trans-activation.82,89,90 This C-terminal region is not associated with changes in levels of HIF-1α protein. Therefore, hypoxia must activate this C-terminal domain by some form of posttranslational modification such as phosphorylation.

HIF-1α has been shown to interact with the transcriptional coactivators, p300 and CBP,92 highly homologous proteins that are functionally and immunologically indistinguishable.93,94 Expression of the adenovirus protein, E1A, which binds to p300/CBP blocking functional activity, prevents hypoxic induction of the genes encoding Epo, VEGF,92 and LDH-A.95 Functional activity of the C-terminal transactivation domain of HIF-1α requires interactions with the CH1 domain of p300/CBP.96,97 p35srj, an alternatively spliced form of MRG-1,98 also binds to the CH1 domain of p300/CBP, competing with HIF-1α for binding to p300/CBP.97 Induction of p35srj by hypoxia may contribute to a negative feedback on HIF-1 activation97 and perhaps explains the prompt decrease in Epo expression after the peak mRNA levels 4 to 6 hours after hypoxic induction.24 p300 and CBP are large proteins with several domains for interacting with multiple proteins. P300/CBP has been shown to interact with HNF-499and other nuclear hormone receptors,100,101CREB,102 TATA binding protein,103 and TFIIB.104 By simultaneously interacting with HIF-1, adjacent transcription factors, and the basal transcriptional machinery, p300/CBP likely acts as a scaffold for the construction of a transcriptionally active, hypoxically inducible complex. Formation of such a complex may explain the requirement of an HNF-4 site, adjacent to the HIF-1 site, for hypoxically inducible activity of the Epo3′ enhancer.95 

A plethora of additional members of the bHLH-PAS family have been cloned recently, including ARNT2, HIF-2α, MOP3, MOP4, and CLOCK.105-109 HIF-2α (also known as EPAS-1, MOP2, HRF, and HLF) has a broad tissue distribution and, like HIF-1α, accumulates only under hypoxic conditions.110 MOP3-HIF-1α and MOP3-HIF-2α heterodimers are capable of binding an HIF consensus site and activating transcription in response to hypoxia.111 ARNT is capable of forming homodimers as well as heterodimers with HIF-1α, AHR, and SIM. The requirement of ARNT for the responses to both hypoxia and aryl hydrocarbons can lead to a functional interference between the 2 pathways.81 

Insulin and insulin-like growth factor I (IGF-1) activate HIF-1 DNA-binding activity, HIF-1–mediated transcriptinal activation of a reporter gene,112 and expression of several oxygen regulated genes, including Epo.113 The signalling pathways for hypoxia, insulin, and IGF-1 appear to converge, all leading to stabilization of HIF-1α.

Activition of HIF-1 is modulated by a complex set of mechanisms likely to include not only protein stability, but also phosphorylation,114,115 redox chemistry,86,87,116 and nuclear localization.96Subsequent to binding of HIF-1 to its cognate cis-acting sequence, interaction with adjacent transcription factors and coactivator proteins are necessary for hypoxic induction of transcription.

Despite a great deal of attention and experimental data, the precise nature of the mammalian oxygen sensor remains elusive. However, the synthesis of evidence obtained to date provides an outline of the likely characteristics of the oxygen sensor that signals the activation of gene expression via HIF-1. A plausible model is shown in Fig 4.

Fig. 4.

Proposed model of oxygen sensing and signaling. In oxygenated cells, a flavo-heme protein functions as an NADPH oxidase, transferring electrons through the flavin (FAD) and heme to molecular oxygen, generating superoxide (O2), which, in the presence of iron, is converted to hydroxyl radical (OH·) and other reactive oxygen species (ROS). As a result, HIF-1 is oxidatively modified so that it is recognized by the proteasome and rapidly degraded. Cobalt (Co2+) as well as other transition metals (Ni2+ and Mn2+) may block the iron-dependent degradation of HIF-1. At low oxygen tension, as well as in the presence of an iron chelator or one of the above-mentioned transition metals, HIF-1 is stable and can form a heterodimer with constitutively expressed HIF-1β, thereby activating HIF-1, which translocates to the nucleus and binds to response elements in hypoxia inducible genes.

Fig. 4.

Proposed model of oxygen sensing and signaling. In oxygenated cells, a flavo-heme protein functions as an NADPH oxidase, transferring electrons through the flavin (FAD) and heme to molecular oxygen, generating superoxide (O2), which, in the presence of iron, is converted to hydroxyl radical (OH·) and other reactive oxygen species (ROS). As a result, HIF-1 is oxidatively modified so that it is recognized by the proteasome and rapidly degraded. Cobalt (Co2+) as well as other transition metals (Ni2+ and Mn2+) may block the iron-dependent degradation of HIF-1. At low oxygen tension, as well as in the presence of an iron chelator or one of the above-mentioned transition metals, HIF-1 is stable and can form a heterodimer with constitutively expressed HIF-1β, thereby activating HIF-1, which translocates to the nucleus and binds to response elements in hypoxia inducible genes.

Close modal

Oxygen sensors in bacteria and yeast have been well characterized, and in these systems heme proteins play a central role.117Similarly, several lines of evidence indicate that a heme protein is involved in mammalian oxygen sensing.118 CO binds specifically and noncovalently to heme proteins, causing the heme moiety to be maintained in an “oxy” conformation. CO blocks the activation of HIF-1 by hypoxia and the hypoxically regulated expression of VEGF, PDGF, ET-1, and PEPCK.117 CO binds with a lower affinity than oxygen to the putative heme protein sensor, with a Haldane coefficient of approximately 0.5.91 The transition metals Co2+, Ni2+, and Mn2+ mimic the effect of hypoxia on Epo, HIF-1, and other oxygen-regulated genes.118,119 A possible mechanism for the action of these transition metals is that they can substitute for iron in heme proteins, locking the heme moiety in the “deoxy” conformation.

Many experiments have indicated that a decrease in levels of oxygen free radicals after hypoxic stimulus leads to the accumulation of HIF-1α and the activation of HIF-1. Addition of H2O2 or agents that increase intracellular peroxide concentration block the induction of Epo and the accumulation of HIF-1α.86 Desferrioxamine (DFO), an iron chelator, mimics the effect of hypoxia on oxygen-regulated genes and HIF-1 activity.120 Intracellular iron probably functions as a Fenton reagent, catalyzing the formation of reactive oxygen species. Therefore, iron chelation, like hypoxia, is likely to effect a reduction in intracellular levels of hydroxyl radical and singlet oxygen. A substantial number of enzymes are inactivated by oxygen-dependent and iron-dependent oxidation at specific residues, rendering them targets for proteolytic degradation.121,122This modification is inhibited by Mn2+ and also by Co2+ and Ni2+ (W. Willmore, R. Levine, and H.F. Bunn, unpublished observations). If this degradative pathway applies to HIF-1α (as shown in Fig 4), it would explain the stabilization of HIF-1 by these 3 transition metals.

Spectrophotometric evidence points to the involvement of a cytochromeb-like protein in oxygen sensing.123-126Furthermore, diphenyl iodonium (DPI), which inhibits NAD(P)H oxidases and other flavoproteins, impairs oxygen sensing.127,128 DPI also inhibits the response to hypoxia in the carotid body129 and pulmonary neuroepithelial bodies.130 However, it is unlikely that the oxygen sensor is identical to the NAD(P)H oxidase in neutrophils and macrophages, which is dedicated to the oxidative burst, necessary for the destruction of engulfed microorganisms.131 Patients with chronic granulomatous disease, who have a genetic defect in 1 of the 4 subunits of neutrophil/macrophage NAD(P)H oxidase, do not have phenotypic evidence of disordered oxygen sensing. The oxygen sensor must be present in a wide range of tissues, and the generation of free radicals involved in the signaling process is likely to be within a specific cell compartment and highly regulated by tissue oxygen tension. (Many experiments have indicated that the oxygen-sensing pathway that regulates Epo is unaffected by inhibitors of mitochondrial respiration.49,132-134 However, Schumacker et al135 have recently presented evidence suggesting that mitochondrial cytochrome oxidase [complex IV] serves as an oxygen sensor in hepatocytes and cardiac myocytes. Their evidence is based on measurements of HIF-1 activation in Hep3B cells treated with mitochondrial inhibitors and in ρο Hep3B cells lacking functional mitochondria.)

In the nitrogen-fixing bacteria, Rhizobium, oxygen-regulated gene expression is mediated by hypoxically inducible phosphorylation of a transcription factor.136,137 Phosphorylation may play a role in signaling the hypoxic stimulus and regulation of HIF-1 activity in mammalian cells as well. Treatment of nuclear extracts from hypoxic cells with alkaline phosphatase abolishes HIF-1 DNA-binding activity.138 Inhibitors of both serine/threonine and tyrosine kinases block activation of HIF-1.114 Multiple compounds that interfere with phosphorylation cascades have complex effects on HIF-1 activation and oxygen-regulated gene expression,115 but the role of a particular phosphorylation pathway in the activation of HIF-1 has not been conclusively proven. The role of src kinase was proposed to play a critical role in the hypoxia signaling pathway.139 Although subsequent experiments indicate that c-src is not necessary for activation of HIF-1,128,140 expression of v-src increases HIF-1 expression in both normoxia and hypoxia.140 In the HIF-1α protein, mutation of phosphoacceptor sites in the hypoxically inducible domain between amino acids 549-672 did not have a major influence on the magnitude hypoxic induction.89 

In the current model of oxygen sensing, the preponderance of evidence supports the role of a heme protein, likely a cytochrome b-like protein, which signals a decrease in oxygen tension by a decrease in the levels of free radicals. Many gaps in knowledge and areas of conflicting data await future elucidation of the oxygen sensing mechanism responsible for the activation of HIF-1 and the regulation of Epo and other oxygen-responsive genes.

The Epo gene provided an apt model system for identification of HIF-1 and investigation of the mechanism of oxygen sensing. An understanding of oxygen-regulated expression of other genes has provided insight into diverse areas of physiology. Genes shown to be regulated by hypoxia are listed in Table 2.

Table 2.

Genes Induced by Hypoxia

Co2+ DFO HIF-1
Erythropoietin  Yes Yes  
VEGF  Yes  Yes  
PDGF-B  Yes  ND  
TGF-β1 Yes  ND  
PLGF  Yes  ND  
Glucose transporter 1 (Glut-1)  Yes  Yes  
Glucose transporter 2 (Glut-2)  Yes ND  
Glucose transporter 3 (Glut-3)  Yes  ND 
Phosphofructokinase L (PFK-L)  Yes  Yes 
Phosphofructokinase C (PFK-C)  Yes  ND  
Aldolase A (ALD-A) Yes  Yes  
Aldolase C (ALD-C)  Yes  ND 
Triosephosphate isomerase (TPI)  ND  ND 
Glyceraldehyde-3-phosphate dehydrogenase (GAPDH)  Yes  ND 
Phosphoglycerate kinase 1 (PGK-1)  Yes  Yes  
Enolase-1 Yes  Yes  
Pyruvate kinase M  Yes  ND  
Lactate dehydrogenase A (LDH-A)  Yes  Yes  
Phosphoenolpyruvate carboxykinase (PEPCK)  Yes  ND  
Mitochondrially encoded genes No  No  
Tyrosine hydroxylase  Yes  Yes  
Endothelin 1 Yes  ND  
Heme oxygenase  Yes  Yes  
Nitric oxide synthase  Yes  Yes  
Transferrin  Yes  Yes 
Retrotransposon VL30  Yes  Yes  
Tissue factor  Yes Yes  
Adenylate kinase 3  Yes  Yes  
c-fos  Yes  ND 
c-jun  Yes  ND  
p35srj  Yes  Yes 
Co2+ DFO HIF-1
Erythropoietin  Yes Yes  
VEGF  Yes  Yes  
PDGF-B  Yes  ND  
TGF-β1 Yes  ND  
PLGF  Yes  ND  
Glucose transporter 1 (Glut-1)  Yes  Yes  
Glucose transporter 2 (Glut-2)  Yes ND  
Glucose transporter 3 (Glut-3)  Yes  ND 
Phosphofructokinase L (PFK-L)  Yes  Yes 
Phosphofructokinase C (PFK-C)  Yes  ND  
Aldolase A (ALD-A) Yes  Yes  
Aldolase C (ALD-C)  Yes  ND 
Triosephosphate isomerase (TPI)  ND  ND 
Glyceraldehyde-3-phosphate dehydrogenase (GAPDH)  Yes  ND 
Phosphoglycerate kinase 1 (PGK-1)  Yes  Yes  
Enolase-1 Yes  Yes  
Pyruvate kinase M  Yes  ND  
Lactate dehydrogenase A (LDH-A)  Yes  Yes  
Phosphoenolpyruvate carboxykinase (PEPCK)  Yes  ND  
Mitochondrially encoded genes No  No  
Tyrosine hydroxylase  Yes  Yes  
Endothelin 1 Yes  ND  
Heme oxygenase  Yes  Yes  
Nitric oxide synthase  Yes  Yes  
Transferrin  Yes  Yes 
Retrotransposon VL30  Yes  Yes  
Tissue factor  Yes Yes  
Adenylate kinase 3  Yes  Yes  
c-fos  Yes  ND 
c-jun  Yes  ND  
p35srj  Yes  Yes 

The first set of genes encode growth factors, the second set of genes encode enzymes or transporters involved in glucose metabolism, and the third set of genes encode proteins with a wide range of functions. The middle column indicates whether the gene has been shown to be induced by cobalt and/or DFO, thereby implicating the same or a similar mechanism of oxygen sensing. The column on the right indicates whether HIF-1 involvement has been demonstrated either by identification of a functional HIF-1 site or by the lack of hypoxic induction in ARNT deficient cells. Expression of the PLGF and Glut-2 genes is decreased by hypoxia. Induction of the PEPCK gene by glucagon is blocked by hypoxia.

Abbreviation: ND, not determined.

Hypoxia activates angiogenesis by inducing the genes encoding VEGF and other growth factors with angiogenic properties.119,141-145HIF-1–mediated regulation of blood vessel growth appears to be of critical importance, because HIF-1α−/− and ARNT−/− knockout mice do not live beyond embryonic day 8.5 to 9.5, likely owing to a failure of vascular development.146-148 In neoplasms, angiogenesis is essential for tumor growth and metastasis. Solid tumor xenografts composed of ARNT-deficient cells have reduced VEGF expression and grow more slowly than xenografts that can form an intact HIF-1 heterodimer.149 

When oxygen is limited, the rate of glycolysis increases to compensate for a decrease in ATP production via mitochondrial respiration. Long-term adaptation to hypoxia involves the regulation of genes encoding proteins involved in energy metabolism. An increase in glucose uptake in hypoxia is associated with increased expression of the genes encoding 2 glucose transporters, Glut-1 and Glut-3.150,151Genes encoding specific isoenzymes for most if not all steps in the glycolytic pathway are upregulated by hypoxia.59,134,150,152-157 Hypoxia blocks glucagon induction of PEPCK, the rate-limiting gluconeogenic enzyme.158-160 Expression of mitochondrially encoded genes is suppressed by hypoxia,154 but the mechanism of regulation of these genes appears to be independent of the signaling pathway that regulates nuclear-encoded genes and HIF-1.150 

Hypoxic induction of the tyrosine hydroxylase gene aids in the regulation of respiration.65,161 HIF-1 has also been implicated in the regulation of genes encoding type II (inducible) nitric oxide synthase,162,163 heme oxygenase-1,164 transferrin,165 and retrotransposon VL30.166 Oxygen-regulated expression has been demonstrated for endothelin 1,167-169, c-jun, and c-fos.166 Hypoxic induction of tissue factor and adenylate kinase 3 were identified by differential display polymerase chain reaction.151 HIF-1α is involved with increasing apoptosis and decreasing cellular proliferation in response to hypoxia, as was demonstrated by use of embryonic stem cells lacking functional HIF-1α (HIF-1α−/−).170 

The importance of HIF-1 in oxygen-regulated gene expression has been examined in several cell lines, including an ARNT-deficient cell line,85 an HIF-1α–defective cell line created by mutagenesis and selection,171 and embryonic stem cells from HIF-1α147,148,170 and ARNT knockout mice.146None of these cells produce Epo, but oxygen-regulated gene expression is disrupted for a number of other genes having functional HIF-1 response elements. For example, induction of PGK-1 and LDH-A by hypoxia is abolished in ARNT-deficient cells. For other genes, some hypoxic induction is retained in the mutant cell lines, for example, heme oxygenase-1 in HIF-1α mutant cells and VEGF in ARNT-deficient cells. Possible explanations for retained inducibility include alternative dimerization partners and regulation at the level of mRNA stability.

The arrangement of sites in hypoxia-responsive regulatory elements in other genes is similar to that of the Epo 3′ enhancer. For example, in the LDH-A promoter, 3 sites are critical for oxygen regulation in arrangement similar to the tripartite structure of theEpo 3′ enhancer.54 Whereas in the Epogene 3′ enhancer the HIF-1 site is adjacent to an HNF-4 site that is necessary for hypoxic inducibility, in the LDH-A promoter, the HIF-1 site is adjacent to a CREB-1/ATF-1 binding site. In both genes, the necessity of multiple adjacent sites for maximal hypoxia induction is probably due to the formation of a multiprotein complex including p300/CBP.

Comparison of HIF-1 sites characterized in 25 hypoxically inducible genes has resulted in a consensus recognition sequence for HIF-1 DNA-binding: T/GACGTGCGG.172 

In a variety of pathological states, dysregulation of Epo gene expression may cause either anemia or polycythemia.

Anemia is a major complication of most forms of renal failure. Because the anemia of renal failure is due primarily to a decrease in Epo production,173,174 patients are successfully treated by administration of recombinant human Epo.175,176 In mice with diverse forms of renal injury, a decreased number of fibroblast-like interstitial cells express Epo in response to anemia or hypoxia.177 Damage to the kidneys appears to change the threshold for Epo gene expression, but the precise molecular mechanisms have not yet been defined. Renal injury causes an expansion of interstitial cells and an infiltration of CD45+ cells, but the phenotype of the Epo-producing fibroblast-like cells does not appear to change. The sensitivity ofEpo expression to changes in the microenvironment of Epo-producing cells may explain the difficulty of establishing a renal cell culture model of inducible Epo expression.

Inappropriately low levels of erythropoietin have been demonstrated in patients with acquired immunodeficiency syndrome (AIDS),178rheumatoid arthritis and other chronic inflammatory diseases,179 and cancer.180 Inflammatory cytokines have been postulated to play a role in diminishing Epogene expression in these disorders and in the anemia of renal failure. In human hepatoma cell lines and isolated perfused rat kidneys, the inflammatory cytokines tumor necrosis factor-α (TNF-α) and interleukin-1 (IL-1) suppress Epo production.181,182Patients with Itai-itai disease, caused by long-term cadmium intoxication, have inappropriately low Epo levels for their degree of anemia and renal failure.183 

Uremia is the predominant and prototypical clinical syndrome for Epo replacement therapy, but recombinent human Epo (rHuEpo) has been used in a broad range of clinical settings (for review, see Cazzola et al184). For example, rHuEpo is efficacious in the treatment of anemia caused by AZT in human immunodeficiency virus (HIV)-infected patients, chemotherapy for nonmyeloid malignancies, premature birth, cancer, rheumatoid arthritis, and inflammatory bowel disease. The efficacy of rHuEpo is uncertain in patients whose levels of plasma Epo are elevated in keeping with their degree of anemia.

Primary polycythemia is caused by defects of hematopoietic progenitor cells and is associated with low levels of circulating erythropoietin. For example, polycythemia vera (PV) is caused by acquired somatic mutations in hematopoietic stem cells. IGF-1 and angiotensin II may increase proliferation of hematopoietic progenitors and thereby contribute to some forms of polycythemia.185 Familial congenital polycythemias may be caused by erythropoietin receptor mutations. In contrast, secondary polycythemia is generally associated with increased erythropoietin production. Elevated levels of plasma Epo are encountered in systemic hypoxemia, in certain neoplasms, and, less commonly, in disorders that impair oxygen delivery to tissues.

Excessive activation of Epo gene expression can result from impaired oxygen delivery due to high affinity hemoglobin mutants, methemoglobinemias, and 2,3-bisphosphoglycerate deficiency.185 Numerous congenital mutations of both α- and β-globin genes can result in high-affinity hemoglobin molecules. Such patients are often asymptomatic, because impaired oxygen delivery is balanced by polycythemia. Congenital methemoglobinemias also cause an adaptive erythrocytosis, because the buildup of ferri-hemes increases the oxygen affinity of the remaining ferro-hemes. 2,3-bisphosphoglycerate is an allosteric regulator of hemoglobin. An enzymatic defect leading to decreased synthesis of 2,3-bisphosphoglycerate is a rare cause of congenital polycythemia.186 

Chronic arterial hypoxemia often leads to an upregulation of Epo expression, causing a maladaptive erythrocytosis. Patients with chronic obstructive pulmonary disease can develop erythrocytosis, which increases the risk of cor pulmonale. Similarly, patients with right to left cardiac shunts can have extremely high hematocrit levels.

Specific types of neoplasms can also cause overproduction of Epo. Elevated Epo levels are found most commonly in patients with renal carcinomas, Wilms tumor, hepatomas, and cerebellar hemangioblastomas,5 all anatomic sites in which Epo is normally expressed at low levels; less frequently, these tumors cause erythrocytosis. Benign renal tumors can also cause erythrocytosis, possibly due to local ischemia of renal Epo-producing cells.

Mutations in the von Hippel Lindau (VHL) gene are associated with renal and central nervous system carcinomas, both highly vascular tumors that overexpress VEGF. In cells in which the von Hippel Lindau protein is inactivated, the hypoxically inducible genes VEGF, Glut-1, and PDGF-B are expressed at high levels under both normoxic and hypoxic conditions. Introduction of wild-type pVHL into these cells causes expression of these genes to revert to the normal hypoxically inducible pattern by suppressing normoxic expression.187,188 

A congenital and familial polycythemia of unknown etiology has been characterized in 103 patients from Chuvashia, a region in the Russian Federation.189 This condition is characterized by an autosomal recessive pattern of inheritance, high hemoglobin levels (mean, 23 g/dL), high hematocrit levels (mean, 67%), elevated Epo levels, and morbidity and mortality secondary to erythrocytosis, including fatal thrombotic and hemorrhagic complications. The polycythemia is not due to high-affinity hemoglobin, methemoglobinemia, 2,3-bisphosphoglycerate deficiency, or systemic hypoxia. Furthermore, the genetic mutation is not linked to either the Epo or Epo receptor genes. The disorder may therefore be caused by an abnormality in the oxygen sensing-signaling pathway or in a trans-acting factor involved in the regulation of Epo gene expression.

Regulation of erythropoiesis and red blood cell mass relies on modulating Epo gene expression in response to tissue oxygen tension. Developmental, tissue-specific, and environmental signals all contribute to the precise regulation of the Epo gene. Epo production and gene expression is restricted to specific subsets of cells: interstitial fibroblast-like cells in the kidney and, in the liver, Ito cells as well as a subset of hepatocytes. Experiments with transgenic mice have broadly mapped the cis-acting sequences responsible for tissue-specific expression.

The Epo gene has been a model for the regulation of gene expression by oxygen tension. The magnitude of hypoxically inducible transcription of the Epo gene is greater than any other gene known to be regulated by oxygen tension. Human hepatoma cell lines have provided a useful model system for studying inducible expression of theEpo gene. Regulatory sequences in the Epo gene have been dissected and characterized in more detail than any other oxygen-regulated gene. In a plausible model, a cytochromeb-like flavoheme protein senses oxygen tension and regulates production of oxygen free radicals.190 In hypoxia, rapid degradation of HIF-1α by the ubiquitin-proteasome pathway is prevented, leading to the formation of HIF-1α/ARNT heterodimers. These heterodimers bind to HIF-1 sites, interact with adjacent DNA-binding proteins and p300/CBP, and activate gene expression.

The Epo gene has been the portal through which a generalized system of oxygen-regulated gene expression was first identified and described. This mode of molecular adaptation is more fundamental than the regulation of Epo in mammals. An HIF-1–like hypoxically inducible DNA-binding activity was identified in Drosophila melanogaster.191 In mammals, HIF-1 and the oxygen-sensing mechanism that regulates Epo are critical for the regulation of genes involved in angiogenesis, energy metabolism, respiration, vascular tone, and many other processes.

In humans, regulation of the Epo gene provides an elegant and precise mechanism for adjusting red blood cell mass to perturbations in tissue oxygen tension. A more complete understanding of the molecular mechanisms governing induction of the Epo gene may lead to new therapeutic agents to treat patients with anemia or polycythemia due to inappropriate expression of the Epo gene.

The authors thank E. Huang, W. Willmore, P. Hradecky, P. Yachimski, and M. Vasconcelles for critical review of the manuscript.

Supported by a National Institutes of Health Grant No. DK41234 to H.F.B.

1
Jourdanet
D
De l’anemie des altitudes et de l’anemie en general dans ses rapports avec la pression de l’atmosphere.
1863
Balliere
Paris, France
2
Lin
F-K
Suggs
S
Lin
C-H
Browne
JK
Smalling
R
Egrie
JC
Chen
KK
Fox
GM
Martin
F
Stabinsky
Z
Badrawi
SM
Lai
P-H
Goldwasser
E
Cloning and expression of the human erythropoietin gene.
Proc Natl Acad Sci USA
82
1985
7580
3
Jacobs
K
Shoemaker
C
Rudersdorf
R
Neill
SD
Kaufman
RJ
Mufson
A
Seehra
J
Jones
SS
Hewick
R
Fritsch
EF
Kawakita
M
Shimizu
T
Miyake
T
Isolation and characterization of genomic and cDNA clones of human erythropoietin.
Nature
313
1985
806
4
Krantz
SB
Erythropoietin.
Blood
77
1991
419
5
Jelkmann
W
Erythropoietin: Structure, control of production, and function.
Physiol Rev
72
1992
449
6
Jacobsen
LO
Goldwasser
E
Fried
W
Plzak
L
Role of the kidney in erythopoiesis.
Nature
179
1957
633
7
Erslev
AJ
Caro
J
Besarab
A
Why the kidney?
Nephron
41
1985
213
8
Schuster
SJ
Badiavas
EV
Costa-Giomi
P
Weinman
R
Erslev
AJ
Caro
J
Stimulation of erythropoietin gene transcription during hypoxia and cobalt exposure.
Blood
73
1989
13
9
Koury
ST
Bondurant
MC
Koury
MJ
Localization of erythropoietin synthesizing cells in murine kidneys by in situ hybridization.
Blood
71
1988
524
10
Koury
ST
Bondurant
MC
Koury
MJ
Semenza
GL
Localization of cells producing erythropoietin in murine liver by in situ hybridization.
Blood
77
1991
2497
11
Lacombe
C
Silva
J-LD
Bruneval
P
Fournier
J-G
Wendling
F
Casadevall
N
Camilleri
J-P
Bariety
J
Varet
B
Tambourin
P
Pertubular cells are the site of erythropoietin synthesis in the murine hypoxic kidney.
J Clin Invest
81
1988
620
12
Koury
ST
Koury
MJ
Bondurant
MC
Caro
J
Graber
SE
Quantitation of erythropoietin-producing cells in kidneys of mice by in situ hybridization: Correlation with hematocrit, renal erythopoietin mRNA, and serum erythropoietin concentration.
Blood
74
1989
645
13
Maxwell
AP
Lappin
TRJ
Johnston
CF
Bridges
JM
McGeown
MG
Erythropoietin production in kidney tubular cells.
Br J Haematol
74
1990
535
14
Loya
F
Yang
Y
Lin
H
Goldwasser
E
Albitar
M
Transgenic mice carrying the erythropoietin gene promoter linked to lacZ express the reporter in proximal convoluted tubule cells after hypoxia.
Blood
84
1994
1831
15
DaSilva
J
Lacombe
C
Bruneval
P
Casadevall
N
Leporrier
M
Camillieri
J
Bariety
J
Tambourin
P
Varet
B
Tumor cells are the site of erythropoietin synthesis in human renal cancers associated with polycythemia.
Blood
75
1990
577
16
Bachmann
S
Hir
ML
Eckardt
KU
Co-localization of erythopoietin mRNA and ecto-5′-nucleotidase immunoreactivity in peritubular cells of rat renal cortex indicates that fibroblasts produce erythropoietin.
J Histochem Cytochem
41
1993
335
17
Maxwell
PH
Osmond
MK
Pugh
CW
Heryet
A
Nicholls
LG
Tan
CC
Doe
BG
Ferguson
DJP
Johnson
MH
Ratcliffe
PJ
Identification of the renal erythropoietin-producing cells using transgenic mice.
Kidney Int
44
1993
1149
18
Zanjani
ED
Poster
J
Burlington
H
Mann
LI
Wasserman
LR
Liver as the site of Epo formation in the fetus.
J Lab Clin Med
89
1977
640
19
Dame
C
Fahnenstich
H
Hofmann
PF
Abdul-Nour
T
Bartmann
P
Fandrey
J
Erythropoietin mRNA expression in human fetal and neonatal tissues.
Blood
92
1998
3218
20
Schuster
SJ
Koury
ST
Bohrer
M
Salceda
S
Caro
J
Cellular sites of extrarenal and renal erythropoietin production in anemic rats.
Br J Hematol
81
1992
153
21
Maxwell
PH
Ferguson
DJP
Osmond
MK
Pugh
CW
Heryet
A
Doe
BG
Johnson
MH
Ratcliffe
PJ
Expression of a homologously recombined erythopoietin-SV40 T antigen fusion gene in mouse liver: Evidence for erythropoietin production by Ito cells.
Blood
84
1994
1823
22
Ratcliffe
PJ
Jones
RW
Phillips
RE
Nicholls
LG
Bell
JI
Oxygen-dependent modulations of erythropoietin mRNA levels in isolated rat kidneys studied by RNase protection.
J Exp Med
172
1990
657
23
Tan
CC
Eckardt
K-U
Ratcliffe
PJ
Organ distribution of erythopoietin messenger RNA in normal and uremic rats.
Kidney Int
40
1991
69
24
Fandrey
J
Bunn
HF
In vivo and in vitro regulation of erythropoietin mRNA: Measurement by competitive polymerase chain reaction.
Blood
81
1993
617
25
Masuda
S
Okano
M
Yamagishi
K
Nagao
M
Ueda
M
Sasaki
R
A novel site of erythropoietin production.
J Biol Chem
1994
1994
19488
26
Digicaylioglu
M
Bichet
S
Marti
HH
Wenger
RH
Rivas
LA
Bauer
C
Gassmann
M
Localization of specific erythropoietin binding sites in defined areas of the mouse brain.
Proc Natl Acad Sci USA
92
1995
3717
27
Marti
HH
Wenger
RH
Rivas
LA
Straumann
U
Digicaylioglu
M
Henn
V
Yonekawa
Y
Bauer
C
Gassmann
M
Erythropoietin gene expression in human, monkey and murine brain.
Eur J Neurosci
8
1996
666
28
Sakanaka
M
Wen
TC
Matsuda
S
Masuda
S
Morishita
E
Nagao
M
Sasaki
R
In vivo evidence that erythropoietin protects neurons from ischemic damage.
Proc Natl Acad Sci USA
95
1998
4635
29
Hermine
O
Beru
N
Pech
N
Goldwasser
E
An autocrine role for erythropoietin in mouse hematopoietic cell differentiation.
Blood
78
1991
2253
30
Stopka
T
Zivny
JH
Stopkova
P
Prchal
JF
Prchal
JT
Human hematopoietic progenitors express erythropoietin.
Blood
91
1998
3766
31
Kurtz
A
Jelkmann
W
Sinowatz
F
Bauer
C
Renal mesangial cell cultures as a model for study of erythopoietin production.
Proc Natl Acad Sci USA
80
1983
4008
32
Sherwood
JB
Shouval
D
Continuous production of erythropoietin by an established human renal carcinoma cell line: Development of the cell line.
Proc Natl Acad Sci USA
83
1986
165
33
Okabe
T
Urabe
A
Kato
T
Chiba
S
Takaku
F
Production of erythropoietin-like activity by human renal and hepatic carcinomas in cell culture.
Cancer
55
1985
1918
34
Goldberg
MA
Glass
GA
Cunningham
JM
Bunn
HF
The regulated expression of erythropoietin by two human hepatoma cell lines.
Proc Natl Acad Sci USA
84
1987
7972
35
Bondurant
M
Koury
M
Anemia induces accumulation of erythopoietin mRNA in the kidney and liver.
Mol Cell Biol
6
1986
2731
36
Beru
N
McDonald
J
Lacombe
C
Goldwasser
E
Expression of the erythropoietin gene.
Mol Cell Biol
6
1986
2571
37
Goldberg
M
Gaut
CC
Bunn
HF
Erythropoietin mRNA levels are governed by both the rate of gene transcrciption and post-transcriptional events.
Blood
77
1991
271
38
Costa-Giomi
P
Caro
J
Weinmann
R
Enhancement by hypoxia of human erythropoietin gene transcription in vitro.
J Biol Chem
265
1990
10185
39
McDonald
JD
Lin
F-K
Goldwasser
E
Cloning, sequencing, and evolutionary analysis of the mouse erythropoietin gene.
Mol Cell Biol
6
1986
842
40
Shoemaker
CB
Mitsock
LD
Murine erythropoietin gene: Cloning, expression, and human gene homology.
Mol Cell Biol
6
1986
849
41
Galson
DL
Tan
CC
Ratcliffe
PJ
Bunn
HF
Comparison of the human and mouse erythropoietin genes reveals extensive homology in the flanking regions.
Blood
82
1993
3321
42
Ho
V
Acquaviva
A
Duh
E
Bunn
HF
Use of a marked erythropoietin gene for investigation of its cis-acting elements.
J Biol Chem
270
1995
10084
43
Semenza
GL
Trystman
M
Gearhart
JD
Antonarakis
S
Polycythemia in transgenic mice expressing the human erythropoietin gene.
Proc Natl Acad Sci USA
86
1989
2301
44
Madan
A
Lin
C
Hatch
SL
Curtin
PT
Regulated basal, inducible, and tissue-specific human erythropoietin gene expression in transgenic mice requires multiple cis DNA sequences.
Blood
85
1995
2735
45
Semenza
GL
Dureza
RC
Traystman
MD
Gearhart
JD
Antonarakis
SE
Human erythropoietin gene expression in transgenic mice: Multiple transcription initiation sites and cis-acting regulatory elements.
Mol Cell Biol
10
1990
930
46
Semenza
GL
Koury
ST
Nejfelt
MK
Gearhart
JD
Antonarakis
SE
Cell-type-specific and hypoxia-inducible expression of the human erythropoietin gene in transgenic mice.
Proc Natl Acad Sci USA
88
1991
8725
47
Semenza
GL
Nejfelt
MK
Chi
SM
Antonarakis
SE
Hypoxia-inducible nuclear factors bind to an enhancer element located 3′ to the human erythropoietin gene.
Proc Natl Acad Sci USA
88
1991
5680
48
Beck
I
Ramirez
S
Weinmann
R
Caro
J
Enhancer element at the 3′-flanking region controls transcriptional response to hypoxia in the human erythropoietin gene.
J Biol Chem
266
1991
15563
49
Pugh
CW
Tan
CC
Jones
RW
Ratcliffe
PJ
Functional analysis of an oxygen-regulated transcriptional enhancer lying 3′ to the mouse erythropoietin gene.
Proc Natl Acad Sci USA
88
1991
10553
50
Blanchard
KL
Acquaviva
AM
Galson
DL
Bunn
HF
Hypoxic induction of the human erythropoietin gene: Cooperation between the promoter and enhancer, each of which contains steroid receptor response elements.
Mol Cell Biol
12
1992
5373
51
Semenza
GL
Wang
GL
A nuclear factor induced by hypoxia via de novo protein synthesis binds to the human erythropoietin gene enhancer at a site required for transcriptional activation.
Mol Cell Biol
12
1992
5447
52
Pugh
CW
Ebert
BL
Ebrahim
O
Ratcliffe
PJ
Characterisation of functional domains within the mouse erythropoietin 3′ enhancer conveying oxygen-regulated responses in different cell lines.
Biochem Biophys Acta
1217
1994
297
53
Kvietikova
I
Wenger
RH
Marti
HH
Gassman
M
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
1995
4542
54
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
1995
21021
55
Hu
B
Wright
E
Campbell
L
Blanchard
KL
In vivo analysis of DNA-protein interactions on the human erythropoietin enhancer.
Mol Cell Biol
17
1997
851
56
Galson
DL
Tsuchiya
T
Tendler
DS
Huang
LE
Ren
Y
Ogura
T
Bunn
HF
The orphan receptor hepatic nuclear factor 4 functions as a transcriptional activator for tissue-specific and hypoxia-specific erythropoetin gene expression and is antagonized by EAR3/COUP-TF1.
Mol Cell Biol
15
1995
2135
57
Imagawa
S
Goldberg
MA
Doweiko
J
Bunn
HF
Regulatory elements of the erythropoietin gene.
Blood
77
1991
278
58
Gupta
M
Goldwasser
E
The role of the near upstream sequence in hypoxic induction of the erythropoietin gene.
Nucleic Acids Res
24
1996
4768
59
Semenza
GL
Roth
PH
Fang
H-M
Wang
GL
Transcriptional regulation of genes encoding glycolytic enzymes by hypoxia-inducible factor 1.
J Biol Chem
269
1994
23757
60
Imagawa
S
Izumi
T
Miura
Y
Positive and negative regulation of the erythropoietin gene.
J Biol Chem
269
1994
9038
61
Aird
WC
Parvin
JD
Sharp
PA
Rosenberg
RD
The interaction of the GATA-binding proteins and basal transcription factors with GATA-containing core promoters.
J Biol Chem
269
1994
883
62
Beru
N
Smith
D
Goldwasser
E
Evidence suggesting negative regulation of the erythropoietin gene by ribonucleoprotein.
J Biol Chem
265
1990
14100
63
Tarumoto
T
Imagawa
S
Ohmine
K
Nagai
T
Suzuki
N
Yamamoto
M
Higuchi
M
Imai
N
Ozawa
K
Novel mechanism of the renal anemia: NG-monomethyl-L-arginine inhibits erythropoietin gene expression by stimulating GATA transcription factors.
Blood
92
1998
197a
(abstr, suppl 1)
64
Yin
H
Blanchard
KL
Methylation of CPGs in the promoter and first intron of the human EPO gene inhibits transcription by two distinct mechanisms.
Blood
92
1998
67a
(abstr, suppl 1)
65
Czyzyk-Krzeska
MF
Furnari
BA
Lawson
EE
Millhorn
DE
Hypoxia increases rate of transcription and stability of tyrosine hydroxylase mRNA in pheochromocytoma (PC12) cells.
J Biol Chem
269
1994
760
66
Czyzyk-Krzeska
MF
Beresh
JE
Characterization of the hypoxia-inducible protein binding site within the pyrimidine-rich tract in the 3′-untranslated region of the tyrosine hydroxylase mRNA.
J Biol Chem
271
1996
3293
67
Levy
AP
Levy
NS
Goldberg
MA
Post-translational regulation of vascular endothelial growth factor by hypoxia.
J Biol Chem
271
1996
2746
68
Levy
NS
Chung
S
Furneaux
H
Levy
AP
Hypoxic stabilization of vascular endothelial growth factor mRNA by the RNA-binding protein HuR.
J Biol Chem
273
1998
6417
69
Rondon
IJ
MacMillan
LA
Beckman
BS
Goldberg
MA
Schneider
T
Bunn
HF
Malter
JS
Hypoxia upregulates the activity of a novel erythopoietin mRNA binding protein.
J Biol Chem
266
1991
16594
70
Rondon
IJ
Scandurro
AB
Wilson
RB
Beckman
BS
Changes in redox affect the activity of erythropoietin RNA binding protein.
FEBS Lett
359
1995
267
71
Scandurro
AB
Rondon
IJ
Wilson
RB
Tenenbaum
SA
Garry
RF
Beckman
BS
Interaction of erythropoietin RNA binding protein with erythropoietin RNA requires an association with heat shock protein 70.
Kidney Int
51
1997
579
72
McGary
EC
Rondon
IJ
Beckman
BS
Post-transcriptional regulation of erythropoietin mRNA stability by erythropoietin mRNA-binding protein.
J Biol Chem
272
1997
8628
73
Scandurro
AB
Beckman
BS
Common proteins bind mRNAs encoding erythropoietin, tyrosine hydroxylase, and vascular endothelial growth factor.
Biochem Biophys Res Commun
246
1998
436
74
Jiang
BH
Semenza
GL
Bauer
C
Marti
HH
Hypoxia-inducible factor 1 levels vary exponentially over a physiologically relevant range of O2 tension.
Am J Physiol
271
1996
C1172
75
Wang
GL
Semenza
GL
Purification and characterization of hypoxia-inducible factor-1.
J Biol Chem
270
1995
1230
76
Wang
GL
Jiang
B-H
Rue
EA
Semenza
GL
Hypoxia-inducible factor 1 is a basic helix-loop-helix-PAS heterodimer regulated by cellular O2 tension.
Proc Natl Acad Sci USA
92
1995
5510
77
Hoffman
EC
Reyes
H
Chu
F-F
Sander
F
Conley
LH
Brooks
BA
Hankinson
O
Cloning of a factor required for activity of the Ah (dioxin) receptor.
Science
252
1991
954
78
Hankinson
O
The aryl hydrocarbon receptor.
Annu Rev Pharmacol Toxicol
35
1995
307
79
Jiang
B-H
Rue
E
Wang
GL
Roe
R
Semenza
GL
Dimerization, DNA binding and transactivation properties of hypoxia-inducible factor 1.
J Biol Chem
271
1996
17771
80
Reisz-Porszasz
S
Probst
M
Fukunaga
B
Hankinson
O
Identification of functional domains of the aryl hydrocarbon receptor nuclear translocator protein (ARNT).
Mol Cell Biol
14
1994
6075
81
Gradin
K
McGuire
J
Wenger
RH
Kvietikova
I
Whitelaw
M
Toftgard
R
Tora
L
Gassman
M
Poellinger
L
Functional interference between hypoxia and dioxin signal transduction pathways: Competition for recruitment of the Arnt transcription factor.
Mol Cell Biol
16
1996
5221
82
Li
H
Ko
HP
Whitlock
JP
Induction of phosphoglycerate kinase 1 gene expression by hypoxia.
J Biol Chem
271
1996
21262
83
Wang
GL
Semenza
GL
General involvement of hypoxia-inducible factor 1 in transcriptional response to hypoxia.
Proc Natl Acad Sci USA
90
1993
4304
84
Maxwell
PH
Pugh
CW
Ratcliffe
PJ
Inducible operation of the erythropoietin 3′ enhancer in multiple cell lines: Evidence for a widespread oxygen-sensing mechanism.
Proc Natl Acad Sci USA
90
1993
2423
85
Wood
SM
Gleadle
JM
Pugh
CW
Hankinson
O
Ratcliffe
PJ
The role of the aryl hydrocarbon receptor nuclear translocator (ARNT) in hypoxic induction of gene expression.
J Biol Chem
269
1996
15117
86
Huang
LE
Arany
Z
Livingston
DM
Bunn
HF
Activation of hypoxia-inducible transcription factor depends primarily upon redox-sensitive stabilization of its α subunit.
J Biol Chem
271
1996
32253
87
Salceda
S
Caro
J
Hypoxia-inducible factor 1α (HIF-1α) protein is rapidly degraded by the ubiquitin-proteasome system under normoxic conditions.
J Biol Chem
272
1997
22642
88
Huang
LE
Gu
J
Schau
M
Bunn
HF
Regulation of hypoxia-inducible factor 1α is mediated by an oxygen-dependent degradation domain via the ubiquitin-proteasome pathway.
Proc Natl Acad Sci USA
95
1998
7987
89
Pugh
CW
O’Rourke
JF
Nagao
M
Gleadle
JM
Ratcliffe
PJ
Activation of hypoxia-inducible factor-1; definition of regulatory domains within the a subunit.
J Biol Chem
272
1997
11205
90
Jiang
B-H
Zheng
JZ
Leung
SW
Roe
R
Semenza
GL
Transactivation and inhibitory domains of hypoxia-inducible factor 1α.
J Biol Chem
272
1997
19253
91
Huang
LE
Willmore
WG
Gu
J
Goldberg
MA
Bunn
HF
Inhibition of HIF-1 activation by carbon monoxide and nitric oxide: Implications for oxygen sensing and signaling.
J Biol Chem
274
1999
9038
92
Arany
Z
Huang
LE
Eckner
R
Bhattacharya
S
Jiang
C
Goldberg
MA
Bunn
HF
Livingston
DM
Participation by the p300/CBP family of proteins in the cellular response to hypoxia.
Proc Natl Acad Sci USA
93
1996
12969
93
Arany
Z
Sellers
WR
Livingston
DM
Eckner
R
E1A-associated p300 and CREB-associated CBP belong to a conserved family of coactivators.
Cell
77
1994
799
94
Lundblad
JR
Kwok
RPS
Laurance
ME
Harter
ML
Goodman
RH
Adenovirus E1A-associated protein p300 as a functional homologue of the transcriptional co-activator CBP.
Nature
374
1995
85
95
Ebert
BL
Bunn
HF
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
1998
4089
96
Kallio
PJ
Okamoto
K
O’Brien
S
Carrero
P
Makino
Y
Tanaka
H
Poellinger
L
Signal transduction in hypoxic cells: Inducible nuclear translocation and recruitment of the CBP/p300 coactivator by the hypoxia-inducible factor-1a.
EMBO J
17
1998
6573
97
Bhattacharya
S
Michels
C
Leung
M-K
Arany
Z
Kung
A
Livingston
D
Functional role of p35srj, a novel p300/CBP binding protein, during transactivation by HIF-1.
Genes Dev
13
1999
64
98
Shioda
T
Fenner
M
Isselbacher
K
msgl, a novel melanocyte-specific gene, encodes a nuclear protein and is associated with pigmentation.
Proc Natl Acad Sci USA
93
1996
12298
99
Yoshida
E
Aratani
S
Itou
H
Miyagishi
M
Takiguchi
M
Osumu
T
Murakami
K
Fukamizu
A
Functional interaction between CBP and HNF4 in trans-activation.
Biochem Biophys Res Commun
241
1997
664
100
Chakravarti
D
LaMorte
VJ
Nelson
MC
Nakajima
T
Schulman
IG
Juguilon
H
Montminy
M
Evans
RM
Role of CBP/p300 in nuclear receptor signaling.
Nature
383
1996
99
101
Kamei
Y
Xu
L
Heinzel
T
Torchia
J
Kurukawa
R
Gloss
B
Lin
S-C
Heyman
R
Rose
DW
Glass
CK
Rosenfeld
MG
A CBP integrator complex mediates trasncriptional activation and AP-1 inhibition by nuclear receptors.
Cell
85
1996
403
102
Chrivia
JC
Kwok
RS
Lamb
N
Hagiwara
M
Montminy
MR
Goodman
RH
Phosphorylated CREB binds specifically to the nuclear protein CBP.
Nature
365
1993
855
103
Yuan
W
Condorelli
G
Caruso
M
Felsani
A
Giordano
A
Human p300 protein is a coactivator for the transcriptional factor MyoD.
J Biol Chem
271
1996
9009
104
Kwok
RPS
Lundblad
JR
Chrivia
JC
Richards
JP
Bachinger
HP
Brennan
RG
Roberts
SGE
Green
MR
Goodman
RH
Nuclear protein CBP is a coactivator for the transcription factor CREB.
Nature
370
1994
223
105
Hirose
K
Morita
M
Ema
M
Mimura
J
Hamada
H
Fujii
H
Saijo
Y
Gotoh
O
Sogawa
K
Fujii-Kuriyama
Y
cDNA cloning and tissue-specific expression of a novel basic helix-loop-helix/PAS factor (Arnt2) with close sequence similarity to the aryl hydrocarbon receptor nuclear translocator (Arnt).
Mol Cell Biol
16
1996
1706
106
Tian
H
McKnight
SL
Russell
DW
Endothelial PAS domain protein 1 (EPAS1), a transcription factor selectively expressed in endothelial cells.
Genes Dev
11
1997
72
107
Hogenesch
JB
Chan
WK
Jackiw
VH
Brown
RC
Gu
Y-Z
Pray-Grant
M
Perdew
GH
Bradfield
CA
Characterization of a subset of the basic-helix-loop-helix-PAS superfamily that interacts with componenets of the dioxin signaling pathway.
J Biol Chem
272
1997
8581
108
Flamme
I
Frolich
T
Reutern
Mv
Kappel
A
Damert
A
Risau
W
HRF, a putative basic helix-loop-helix-PAS-domain transcription factor is closely related to hypoxia-inducible factor-1a and developmentally expressed in blood vessels.
Mech Dev
63
1997
51
109
Ema
M
Taya
S
Yokotani
N
Sogawa
K
Matsuda
Y
Fujii-Kuriyama
Y
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
1997
4273
110
Wiesener
MS
Turley
H
Allen
WE
William
C
Eckhardt
K-U
Talks
KL
Wood
SM
Gatter
KC
Harris
AL
Pugh
CW
Ratcliffe
PJ
Maxwell
PH
Induction of endothelial PAS domain protein-1 by hypoxia: Characterisation and comparison with hypoxia-inducible factor-1 alpha.
Blood
92
1998
2260
111
Hogenesch
JB
Gu
Y-Z
Jain
S
Bradfield
CA
The basic-helix-loop-helix-PAS orphan MOP3 forms transcriptionally active complexes with circadian and hypoxia factors.
Proc Natl Acad Sci USA
95
1998
5474
112
Zelzer
E
Levy
Y
Kahana
C
Shilo
B-Z
Rubinstein
M
Cohen
B
Insulin induces transcription of target genes through the hypoxia-inducible factor HIF-1α/ARNT.
EMBO J
17
1998
5085
113
Masuda
S
Chikuma
M
Sasaki
R
Insulin-like growth factors and insulin stimulate erythropoietin production in primary cultured astrocytes.
Brain Res
746
1997
63
114
Wang
GL
Jiang
B-H
Semenza
GL
Effect of protein kinase and phosphatase inhibitors on expression of hypoxia-inducible factor 1.
Biochem Biophys Res Commun
216
1995
669
115
Salceda
S
Beck
I
Srinivas
V
Caro
J
Complex role of protein phosphorylation in gene activation by hypoxia.
Kidney Int
51
1997
556
116
Wang
GL
Jiang
B-H
Semenza
GL
Effect of altered redox states on expression and DNA-binding activity of hypoxia-inducible factor 1.
Biochem Biophys Res Commun
212
1995
550
117
Bunn
HF
Poyton
RO
Oxygen sensing and molecular adaptation to hypoxia.
Physiol Rev
76
1996
839
118
Goldberg
MA
Dunning
SP
Bunn
HF
Regulation of the erythopoietin gene: Evidence that the oxygen sensor is a heme protein.
Science
242
1988
1412
119
Gleadle
JM
Ebert
BL
Firth
JD
Ratcliffe
PJ
Regulation of angiogenic growth factor expression by hypoxia, transition metals and chelating agents.
Am J Physiol
268
1995
C1362
120
Wang
GL
Semenza
GL
Desferrioxamine induces erythropoietin gene expression and hypoxia-inducible factor 1 DNA-binding activity: Implications for models of hypoxia signal transduction.
Blood
82
1993
3610
121
Stadtman
ER
Oliver
CN
Metal-catalyzed oxidation of proteins.
J Biol Chem
266
1991
2005
122
Stadtman
ER
Protein oxidation and aging.
Science
257
1992
1220
123
Acker
H
Dufau
E
Huber
J
Sylvester
D
Indications to an NADPH oxidase as a possible PO2 sensor in the rat carotid body.
FEBS Lett
256
1989
75
124
Acker
H
Bölling
B
Delpiano
MA
Dufau
E
Gorläch
A
Holtermann
G
The meaning of H2O2 generation in carotid body cells for PO2 chemoreception.
J Auton Nerv Syst
41
1992
41
125
Acker
H
Mechanisms and meaning of cellular oxygen sensing in the organism.
Respir Physiol
95
1994
1
126
Acker
H
Xue
D
Mechanisms of oxygen sensing in the carotid body in comparison to other oxygen sensing cells.
News Physiol Sci
10
1995
211
127
Goldwasser
E
Alibali
P
Gardner
A
Differential inhibition by iodonium compounds of induced erythropoietin expression.
J Biol Chem
270
1995
2628
128
Gleadle
JM
Ebert
BL
Ratcliffe
PJ
Diphenylene iodonium inhibits the induction of erythropoietin and other mammalian genes by hypoxia: Implications for the mechanism of oxygen sensing.
Eur J Biochem
234
1995
92
129
Cross
AR
Henderson
L
Jones
OTG
Delpiano
MA
Hentschel
J
Acker
H
Involvement of an NAD(P)H oxidase as as pO2 sensor protein in the rat carotid body.
Biochem J
272
1990
743
130
Youngson
C
Nurse
C
Yeger
H
Cutz
E
Oxygen sensing in airway chemoreceptors.
Nature
365
1993
153
131
Wenger
RH
Marti
HH
Schuerer-Maly
CC
Kvietikova
I
Bauer
C
Gassman
M
Maly
FE
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
1996
756
132
Necas
E
Thorling
EB
Unresponsiveness of erythropoietin-producing cells to cyanide.
Am J Physiol
222
1972
1187
133
Tan
CC
Ratcliffe
PJ
Effect of inhibitors of oxidative phosphorylation on erythopoietin mRNA in isolated perfused rat kidneys.
Am J Physiol
261
1991
F982
134
Ebert
BL
Firth
JD
Ratcliffe
PJ
Hypoxia and mitochondrial inhibitors regulate expression of glucose transporter-1 via distinct cis-acting sequences.
J Biol Chem
270
1995
29083
135
Chandel
NS
Maltepe
E
Goldwasser
E
Mathieu
CE
Simon
MC
Schumacker
PT
Mitochondrial reactive oxygen species trigger hypoxia-induced transcription.
Proc Natl Acad Sci USA
95
1998
11715
136
David
M
Daveran
ML
Batut
J
Dedieu
A
Domergue
O
Gai
J
Herteg
C
Boistard
P
Kahn
D
Cascade regulation of nif gene expression in Rhizobium meliloti.
Cell
54
1988
671
137
Gilles-Gonzalez
MA
Ditta
GS
Helinski
DR
A haemoprotein with kinase activity encoded by the oxygen sensor of Rhizobium meliloti.
Nature
350
1991
170
138
Wang
GL
Semenza
GL
Characterization of hypoxia-inducible factor 1 and regulation of DNA binding activity by hypoxia.
J Biol Chem
268
1993
21513
139
Mukhopadhyay
D
Tsiokas
L
Zhou
X-M
Foster
D
Brugge
JS
Sukhatme
VP
Hypoxic induction of human vascular endothelial growth factor expression through c-Src expression.
Nature
375
1995
577
140
Jiang
BH
Agani
F
Passaniti
A
Semenza
GL
V-SRC induces expression of hypoxia-inducible factor 1 (HIF-1) and transcription of genes encoding vascular endothelial growth factor and enolase 1: Involvement of HIF-1 in tumor progression.
Cancer Res
57
1997
5328
141
Shweiki
D
Itin
A
Soffer
D
Keshet
E
Vascular endothelial growth factor induced by hypoxia-initiated angiogenesis.
Nature
359
1992
843
142
Goldberg
MA
Schneider
TJ
Similarities between the oxygen-sensing mechanisms regulating the expression of vascular endothelial growth factor and erythropoietin.
J Biol Chem
269
1994
4355
143
Kourembanas
S
Hannan
RL
Faller
DV
Oxygen tension regulates the expression of the platelet-derived growth factor-B chain gene in human endothelial cells.
J Clin Invest
86
1990
670
144
Levy
A
Levy
N
Wegner
S
Goldberg
M
Transcriptional regulation of the rat vascular endothelial growth factor gene by hypoxia.
J Biol Chem
270
1995
13333
145
Forsythe
JA
Jiang
B-H
Iyer
NV
Agani
F
Leung
SW
Koos
RD
Semenza
GL
Activation of vascular endothelial growth factor gene transcription by hypoxia-inducible factor 1.
Mol Cell Biol
16
1996
4604
146
Maltepe
E
Schmidt
JV
Baunoch
D
Bradfield
CA
Simon
MC
Abnormal angiogenesis and responses to glucose and oxygen deprivation in mice lacking the protein ARNT.
Nature
386
1997
403
147
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
1998
149
148
Ryan
HE
Lo
J
Johnson
RS
HIF-1 alpha is required for solid tumor formation and embryonic vascularization.
EMBO J
17
1998
3005
149
Maxwell
PH
Dachs
GU
Gleadle
JM
Nicholls
LG
Harris
AL
Stratford
IJ
Hankinson
O
Pugh
CW
Ratcliffe
PJ
Hypoxia-inducible factor-1 modulates gene expression in solid tumors and influences both angiogenesis and tumor growth.
Proc Natl Acad Sci USA
94
1997
8104
150
Ebert
BL
Gleadle
JM
O’Rourke
JF
Bartlett
SM
Poulton
J
Ratcliffe
PJ
Isoenzyme-specific regulation of genes involved in energy metabolism by hypoxia: Similarities with the regulation of erythropoietin.
Biochem J
313
1996
809
151
O’Rourke
JF
Pugh
CW
Bartlett
SM
Ratcliffe
PJ
Identification of hypoxically inducible mRNAs in HeLa cells using differential-display PCR. Role of hypoxia-inducible factor-1.
Eur J Biochem
241
1996
403
152
Webster
KA
Regulation of glycolytic enzyme RNA transcriptional rates by oxygen availability in skeletal muscle cells.
Mol Cell Biochem
77
1987
19
153
Webster
KA
Murphy
BJ
Regulation of tissue-specific glycolytic isozyme genes: Coordinate response to oxygen availability in myogenic cells.
Can J Zool
66
1988
1046
154
Webster
KA
Gunning
P
Hardeman
E
Wallace
DC
Kedes
L
Coordinate reciprocal trends in glycolytic and mitochondrial transcript accumulations during the in vitro differentiation of human myoblasts.
J Cell Physiol
142
1990
566
155
Firth
JD
Ebert
BL
Pugh
CW
Ratcliffe
PJ
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
1994
6496
156
Semenza
GL
Jiang
B-H
Leung
SW
Passatino
R
Concordet
J-P
Maire
P
Giallongo
A
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
1996
32529
157
Graven
K
McDonald
R
Farber
H
Hypoxic regulation of endothelial glyceraldehyde-3-phosphate dehydrogenase.
Am J Physiol
274
1998
C347
158
Hellkamp
J
Christ
B
Bastian
H
Jungermann
K
Modulation by oxygen of the glucagon-dependent activation of the phosphoenolpyruvate carboxykinase gene in rat hepatocyte cultures.
Eur J Biochem
198
1990
635
159
Kietzmann
T
Schmidt
H
Probst
I
Jungermann
K
Modulation of the glucagon-dependent activation of the phosphoenolpyruvate carboxykinase gene by oxygen in rat hepatocyte cultures.
FEBS Lett
311
1992
251
160
Kietzmann
T
Schmidt
H
Unthan-Fechner
K
Probst
I
Jungermann
K
A ferro-heme protein senses oxygen levels, which modulate the glucagon dependent activation of the phosphoenolpyruvate carboxykinase in rat hepatocyte cultures.
Biochem Biophys Res Commun
195
1993
792
161
Czyzyk-Krzeska
MR
Bayliss
DA
Lawson
EE
Millhorn
DE
Regulation of tyrosine hyroxylase gene expression in the rat carotid body by hypoxia.
J Neurochem
58
1992
1538
162
Melillo
G
Musso
T
Sica
A
Taylor
LS
Cox
GW
Varesio
L
A hypoxia-responsive element mediates a novel pathway of activation of the inducible nitric oxide synthase promoter.
J Exp Med
182
1995
1683
163
Palmer
L
Semenza
G
Stoler
M
Johns
R
Hypoxia induces type II NOS gene expression in pulmonary artery endothelial cells via HIF-1.
Am J Physiol
274
1998
L212
164
Lee
P
Jiang
B
Chin
B
Iyer
N
Alam
J
Semenza
G
Choi
A
Hypoxia-inducible factor-1 mediates transcriptional activation of the heme oxygenase-1 gene in response to hypoxia.
J Biol Chem
272
1997
5375
165
Rolfs
A
Kvietikova
I
Gassmann
M
Wegner
R
Oxygen-regulated transferrin expression is mediated by hypoxia-inducible factor-1.
J Biol Chem
273
1997
20055
166
Estes
S
Stoler
D
Anderson
G
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
1995
6335
167
Firth
J
Ratcliffe
P
Organ distribution of the three rat endothelin messenger RNAs and the effects of ischemia on renal gene expression.
J Clin Invest
90
1992
1023
168
Kourembanas
S
Marsden
PA
McQuillan
LP
Faller
DV
Hypoxia induces endothelin gene expression and secretion in cultured human endothelium.
J Clin Invest
88
1991
1054
169
Bodi
I
Bishopric
N
Discher
D
Wu
X
Webster
K
Cell-specificity and signaling pathway of endothelin-1 gene regulation by hypoxia.
Cardiovasc Res
30
1995
975
170
Carmeliet
P
Dor
Y
Herbert
J-M
Fukumura
D
Brusselmans
K
Dewerchin
M
Neeman
M
Bono
F
Abramovitch
R
Maxwell
P
Koch
CJ
Ratcliffe
P
Moons
L
Jain
RK
Collen
D
Keshet
E
Role of HIF-1α in hypoxia-mediated apoptosis, cell proliferation and tumour angiogenesis.
Nature
394
1998
485
171
Wood
SM
Wiesener
MS
Yeates
KM
Okada
N
Pugh
CW
Maxwell
PH
Ratcliffe
PJ
Selection and analysis of a mutant cell line defective in the hypoxia-inducible factor-1 α-subunit (HIF-1α).
J Biol Chem
273
1998
8360
172
Wenger
RH
Gassman
M
Oxygen(es) and the hypoxia inducible factor-1.
Biol Chem
378
1997
609
173
Caro
J
Brown
S
Miller
O
Murray
T
Erslev
AJ
Erythropoietin levels in uremic nephric and anephric patients.
J Lab Clin Med
93
1979
449
174
Cotes
PM
Physiological studies of erythropoietin in plasma
Erythropoietin.
Jelkmann
W
Gross
AJ
1989
57
Springer
Berlin, Germany
175
Winearls
C
Oliver
D
Pippard
M
Reid
C
Downing
M
Cotes
P
Effect of human erythropoietin derived from recombinant DNA on the anemia of patients maintained on chronic hemodialysis.
Lancet
2
1986
1175
176
Eschbach
J
Ergie
J
Downing
M
Browne
J
Adamson
J
Correction of the anemia of endstage renal disease with recombinant human erythropoietin.
N Engl J Med
316
1987
73
177
Maxwell
PH
Ferguson
DJ
Nicholls
LG
Johnson
MH
Ratcliffe
PJ
The interstitial response to renal injury: fibroblast-like cells show phenotypic changes and have reduced potential for erythropoietin gene expression.
Kidney Int
52
1997
715
178
Fischl
M
Galpin
JE
Levine
JD
Groopman
JE
Henry
DH
Kennedy
P
Miles
S
Robbins
W
Starrett
B
Zalusky
R
Abels
RI
Tsai
HC
Rudnick
SA
Recombinant human erythropoietin for patients with AIDS treated with zidovudine.
N Engl J Med
322
1990
1488
179
Means
JRT
Clinical application of recombinant erythropoietin in the anemia of chronic disease.
Hematol Oncol Clin North Am
8
1994
933
180
Spivak
JL
Recombinant human erythropoietin ns the anemia of cancer.
Blood
84
1994
997
181
Faquin
WC
Schneider
TJ
Goldberg
MA
Effect of inflammatory cytokines on hypoxia-induced erythropoietin production.
Blood
79
1992
1987
182
Jelkmann
W
Pagel
WH
Wolff
M
Fandrey
J
Monokines inhibiting erythropoietin production in human hepatoma cultures and in isolated perfused rat kidneys.
Life Sci
50
1992
301
183
Horiguchi
H
Teranishi
H
Niiya
K
Aoshima
K
Katoh
T
Sakuragawa
N
Kasuya
M
Hypoproduction of erythropoietin contributes to anemia in chronic cadmium intoxication: clinical study on Itai-itai disease in Japan.
Arch Toxicol
68
1994
632
184
Cazzola
M
Mercuriali
F
Brugnara
C
Use of recombinant human erythropoietin outside the setting of uremia.
Blood
89
1997
4248
185
Prchal
JF
Prchal
JT
Molecular basis for polycythemia.
Curr Opin Hematol
6
1999
100
186
Rosa
R
Prehu
M-O
Beuzard
Y
Rosa
J
The first case of a complete deficiency of diphosphoglycerate mutase in human erythrocytes.
J Clin Invest
62
1978
907
187
Gnarra
J
Zhou
S
Merrill
M
Wagner
J
Krumm
A
Papavassiliou
E
Oldfield
E
Klausner
R
Linehan
W
Post-transcriptional regulationof vascular endothelial growth factor mRNA by the product of the VHL tumor suppressor gene.
Proc Natl Acad Sci USA
93
1996
10589
188
Iliopoulos
O
Levy
A
Jiang
C
WG Kaelin
J
Goldberg
M
Negative regulation of hypoxia-inducible genes by the von Hippel-Lindau protein.
Proc Natl Acad Sci USA
93
1996
10595
189
Sergeyeva
A
Gordeuk
VR
Tokarev
YN
Sokol
L
Prchal
JH
Prchal
JT
Congenital polycythemia in Chuvashia.
Blood
89
1997
2148
190
Zhu
H
Bunn
HF
Oxygen sensing and signaling: Impact on the regulation of physiologically important genes.
Resp Physiol
115
1999
239
191
Nagao
M
Ebert
BL
Ratcliffe
PJ
Pugh
CW
Drosophila melanogaster SL2 cells contain a hypoxically inducible DNA binding complex which recognizes mammalian HIF-1 binding sites.
FEBS Lett
387
1996
161
192
Goldwasser
E
Jacobson
LO
Fried
W
Plzak
LF
The effect of cobalt on the production of erythropoietin.
Studies Erythropoiesis V
13
1958
55
193
Eckardt
K-U
Pugh
CW
Ratcliffe
PJ
Kurtz
A
Oxygen dependent modulation of erythropoietin mRNA in rat hepatocytes in vitro.
Pflugers Arch
423
1993
356
194
Fandrey
J
Frede
S
Jelkmann
W
Role of hydrogen peroxide in hypoxia-induced erythropoietin production.
Biochem J
303
1994
507
195
Jelkmann
W
Huwiler
A
Fandrey
J
Pfeilschifter
J
Inhibition of erythropoietin production by phorbol ester is associated with down-regulation of protein kinase C-a isoenzyme in hepatoma cells.
Biochem Biophys Res Commun
179
1991
1441
196
Kurtz
A
Eckartd
K-U
Pugh
C
Corvol
P
Fabbro
D
Ratcliffe
P
Phorbol ester inhibits erythropoietin production in human hepatoma cells.
Am J Physiol
262
1992
C1204
197
Faquin
WC
Schneider
TJ
Goldberg
MA
Modulators of protein kinase C inhibit hypoxia-induced erythropoietin production.
Exp Hematol
21
1993
420

Author notes

Address reprint requests to H. Franklin Bunn, MD, LMRC 223, 221 Longwood Ave, Boston, MA 02115; e-mail: bunn@calvin.bwh.harvard.edu.

Sign in via your Institution