Organisms living under aerobic conditions need oxygen for the metabolic conversion of nutrition into energy. With the appearance of increasingly complex animals, a specialized transport system (erythrocytes) arose during evolution to provide oxygen to virtually every single cell in the body. Moreover, in case of low environmental partial pressure of oxygen, the number of erythrocytes automatically increases to preserve sustained oxygen delivery. This process relies predominantly on the cytokine erythropoietin (Epo) and its transcription factor hypoxia inducible factor (HIF), whereas the von Hippel-Lindau (VHL) ubiquitin ligase as well as the oxygen-sensitive prolyl hydroxylases (PHDs) represent essential regulators of this oxygen-sensing system. Deregulation of particular members of this pathway (eg, PHD2, HIF2α, VHL) lead to disorders in blood homeostasis as a result of insufficient (anemia) or excessive (erythrocytosis) red blood cell production.

High altitude is accompanied by low atmospheric oxygen pressure, which sequentially leads to insufficient oxygen uptake and reduced tissue oxygenation. In general, inadequate oxygen supply is detrimental and might lead to the death of cells, tissues, or, ultimately, even the organism. To avoid this, complex cardiovascular, respiratory, and hematologic mechanisms have evolved, and one such long-term adaptation process is the elevation of erythrocyte numbers to boost the blood’s oxygen transport capacity.

As early as the 19th century, scientists recognized the correlation between low atmospheric oxygen pressure and elevated red blood cell numbers in humans and animals.1  Several decades later, it became evident that low oxygen does not directly act on hematopoietic cells but induces the production of a soluble factor called erythropoietin (Epo). In 1977, Epo was purified from the urine of anemic patients2  and in 1985 the corresponding EPO gene was isolated and cloned.3  Approximately 2 decades ago, the transcription factor hypoxia inducible factor (HIF) was first identified in hepatoma cells as the regulator of Epo through its binding to a hypoxia-responsive element (HRE) present in the 3′ enhancer region of the EPO gene.4,5  In subsequent work, HIF was found to be expressed widespread in mammalian cells and even in lower animals that do not produce Epo or red blood cells.6-8  In the past decennium, much knowledge has been acquired regarding the role of HIFs during red blood cell production (erythropoiesis). Accordingly, this review aims to discuss recent findings on these essential proteins in humans and mice, and the detrimental impact of their deregulation.

Epo, a glycoprotein hormone, is the principal stimulator of erythropoiesis and is induced under hypoxic conditions. In 1957, the kidney was first identified as the primary Epo-producing organ in adult mammals,9  whereas the liver is the major source of Epo during embryogenesis (eg, hepatocytes, Ito cells). In the kidney, a specialized Epo-producing cell (REPC) was identified in the cortex and outer medulla that was initially described as an interstitial fibroblastlike cell with neuronal characteristics.10-14  Interestingly, and in contrast to other organs, the kidney is able to increase the total amount of REPCs in an oxygen-dependent manner rather than increasing Epo expression per cell.11,14  Neurons and glial cells in the central nervous system represent an additional source of Epo;15,16  it has been suggested that the glycohormone functions in a paracrine fashion as a protective,17  ventilation,18  or cognition-enhancing factor.19 

Erythropoiesis is a complex multistep process during which erythroid progenitors enucleate and develop into mature red blood cells. Upon Epo binding to its receptor, the EpoR signaling through the Janus kinase 2 (JAK2) activates multiple pathways including Stat5, phosphoinositide-3 kinase/Akt, and p42/44 mitogen-activated protein kinase. This reduces apoptosis and promotes expansion and differentiation of the progenitors.20  In adult mammals, erythropoiesis is mainly carried out by the bone marrow. However, in response to stress (eg, anemia, bone marrow transplant, certain diseases), erythropoiesis may extend to extramedullary sites, such as the spleen and liver, thereby increasing erythrocyte output. As has been shown in mice, stress erythropoiesis is characterized by massive self-renewal of burst-forming unit–erythroid cells and is regulated by additional extrinsic factors such as the stress hormone cortisol, stem cell factor, and the bone morphogenetic protein 4.21-24  In humans, analogous pathways have not yet been identified, and the molecular basis is also not well-described. However, recently, erythroblastic island macrophages have been reported to facilitate human stress and pathological erythropoiesis.25 

The HIF pathway is present in virtually every cell of the body, orchestrating a whole cascade of downstream genes that allow acclimatization to reduced levels of oxygen. The α subunit of the HIFs (mainly HIF1α and HIF2α) becomes stabilized nearly instantaneously during low oxygen conditions26  and translocates to the nucleus, where it dimerizes with the constitutively expressed HIFβ subunit and promotes transcription of genes containing a HRE.27  In human cells, pan-genomic analyses of HIF binding to DNA have now revealed the existence of >500 direct transcriptional targets of HIF in a given cell line.28,29  More than a decade ago, the groups of Ratcliffe and Kaelin discovered that both HIF1α and HIF2α are regulated at the posttranscriptional level by the HIF prolyl-hydroxylase domain enzymes (PHDs).30,31  These oxygen sensors hydroxylate the α-subunits and prime them for poly-ubiquitination by the von Hippel-Lindau (VHL) tumor suppressor complex, which ultimately leads to proteolytic degradation (Figure 1). To date, 4 PHDs have been identified in mammals, of which PHD2 (gene name: Egl nine homolog 1 (Egln1)) has been described as the key limiting enzyme targeting HIFα for degradation under normoxic and mild hypoxic conditions.32-34 

Figure 1

Oxygen-dependent regulation of HIFα and its target genes. If oxygen demand is covered, HIFα becomes constantly hydroxylated by PHDs and subsequently undergoes proteasomal degradation after binding with VHL. Under hypoxic conditions, HIFα is stabilized, translocates to the nucleus, binds to its heterodimerization partner HIFβ as well as to other co-factors, and leads to the transcriptional activation of target genes that harbor HRE sequences in their promoter region. HIF1α and HIF2α share target genes (green) but have certain preferences (preferentially induced by HIF1α [yellow] and HIF2α [blue]). CBP, CREB-binding protein.

Figure 1

Oxygen-dependent regulation of HIFα and its target genes. If oxygen demand is covered, HIFα becomes constantly hydroxylated by PHDs and subsequently undergoes proteasomal degradation after binding with VHL. Under hypoxic conditions, HIFα is stabilized, translocates to the nucleus, binds to its heterodimerization partner HIFβ as well as to other co-factors, and leads to the transcriptional activation of target genes that harbor HRE sequences in their promoter region. HIF1α and HIF2α share target genes (green) but have certain preferences (preferentially induced by HIF1α [yellow] and HIF2α [blue]). CBP, CREB-binding protein.

Close modal

HIF1α exhibits a ubiquitous expression pattern, whereas HIF2α is found in a limited number of cell types including endothelial cells, cardiomyocytes, hepatocytes, glial cells, and interstitial cells of the kidney.35  Both isoforms have overlapping sets of target genes but can also play nonredundant roles, depending on the cell type and oxygen concentrations.36  Accordingly, HIF1α has been suggested to represent the response to acute hypoxia, whereas HIF2α is the predominant subunit to chronic exposure to low oxygen that occurs at high altitudes.37  In addition, several studies have demonstrated that both HIF isoforms can even display opposing roles in vivo, for example, in renal cell carcinoma growth and metastasis formation.38,39  Glycolysis enzymes like phosphoglycerate kinase 1 and lactate dehydrogenase A are predominantly HIF1α-dependent.36  In contrast, HIF2α has been described to induce matrix metallopeptidase 9 and the transcription factor oct-4, which is involved in stem cell function and the elevation of hemoglobin gene expression in humans (Figure 1).40-42  Until recently, it was unclear which of the HIF and PHD isoforms regulated erythropoiesis, and the expression of Epo in particular. Only with knowledge gained from patients with erythrocytosis and transgenic mice did it become evident that the HIF2α isoform and not HIF1α is the key player in EPO gene expression and erythropoiesis-enhancing processes (eg, iron absorption and transport).

Erythrocytosis is an aberrant increase in red blood cell numbers and comprises a heterogeneous group of disorders. A general distinction is made between the hypersensitivity of the erythroid progenitors to Epo (primary erythrocytosis) and the excessive activation of EPO gene transcription (secondary erythrocytosis). The most common example of primary erythrocytosis is polycythemia vera. Here, erythroid progenitors carry a gain-of-function mutation in the JAK2 gene, which leads to constitutive activation of the EPO signaling pathway at the EPO-R level. On the other hand, patients bearing point mutations in specific members of the HIF pathway can develop secondary erythrocytosis (Table 1).

Table 1

HIF pathway–related mutations that have resulted in erythrocytosis and/or tumor development in humans

GeneType of mutationMutationErythrocytosisTumor typeRef.
VHL Germline (Homo>>Hetero) R200W + — 45,55  
 Germline (Homo) H191D P138L + — 57,-59  
 Germline ((Compound) Hetero) Various (including R200W) + — 55,56  
 Somatic/ Germline (Hetero) Various  E.g. spinal hemangioblastoma, renal cell carcinoma (RCC), and pheochromocytoma 60  
PHD2 Germline (Hetero) Various + — 55,61,,-64,66,67  
 Germline (Hetero) H374R + Paraganglioma 68  
HIF2α Germline (Hetero) Various (including G537W) + — 55,69,70  
 Germline (Hetero) F374Y + Pheochromocytoma/paraganglioma 73  
 Somatic (Hetero) A530V A530T + Paragangliomas/somatostatinoma 74  
 Somatic (Hetero) Various +/− Pheochromocytomas/paragangliomas/somatostatinoma 75,,-78  
GeneType of mutationMutationErythrocytosisTumor typeRef.
VHL Germline (Homo>>Hetero) R200W + — 45,55  
 Germline (Homo) H191D P138L + — 57,-59  
 Germline ((Compound) Hetero) Various (including R200W) + — 55,56  
 Somatic/ Germline (Hetero) Various  E.g. spinal hemangioblastoma, renal cell carcinoma (RCC), and pheochromocytoma 60  
PHD2 Germline (Hetero) Various + — 55,61,,-64,66,67  
 Germline (Hetero) H374R + Paraganglioma 68  
HIF2α Germline (Hetero) Various (including G537W) + — 55,69,70  
 Germline (Hetero) F374Y + Pheochromocytoma/paraganglioma 73  
 Somatic (Hetero) A530V A530T + Paragangliomas/somatostatinoma 74  
 Somatic (Hetero) Various +/− Pheochromocytomas/paragangliomas/somatostatinoma 75,,-78  

In 1997, the first type of erythrocytosis related to the HIF pathway was discovered by Prchal and colleagues. They described 103 individuals from 81 families living in the Chuvash region in Russia with erythrocytosis.43  Several patients were studied in detail and displayed markedly increased hematocrit levels accompanied by significantly higher levels of Epo. However, molecular analysis failed to demonstrate mutations in the EPO-R or previously described erythrocyte alterations (eg, high oxygen affinity hemoglobin). Subsequent genetic studies revealed a homozygous mutation in the VHL gene (C598T leading to the R200W amino acid change) in all affected individuals. This resulted in reduced affinity of VHL for the hydroxylated HIFα subunit and subsequent increase of Epo and red blood cells.44,45  Recently, the underlying molecular mechanism was discovered: the R200W VHL mutation alters the affinity of VHL for suppression of cytokine signaling 1, which prevents the degradation of the EPO-R coupled kinase pJAK2.46  This illustrates that VHL, as part of the oxygen-sensing machinery, does not only influence the production of Epo but also regulates erythropoiesis at different levels. The clinical presentation of patients with Chuvash erythrocytosis has been carefully studied and includes a wide range of hematologic and vascular abnormalities, but no tumors. Patients with Chuvash erythrocytosis have complications such as thrombosis, major bleeding episodes, and higher systolic pulmonary artery pressure, which collectively lead to premature lethality.47-49  However, it has been suggested that thromboembolic events in patients with VHL mutations might be associated with a subsequent gain in HIFα activity rather than the increase in red blood cell mass. For instance, vascular endothelial growth factor and plasminogen activator inhibitor-1, 2 HIFα targets, are upregulated in the serum of patients with Chuvash erythrocytosis and might have an impact on coagulation pathways.50,51  Further studies have revealed higher homocysteine levels in patients with Chuvash erythrocytosis, which could be an additional cause for the observed elevated blood pressure and thrombosis.52  Later, another cohort with the same mutation was identified on an island in the Bay of Naples, Italy, which suggested a founder mutation. Indeed, single nucleotide polymorphism analysis near the VHL gene on persons from different ethnic backgrounds confirmed this, indicating that the R200W mutation arose between 14 000 and 62 000 years ago in a single ancestor.53,54  Apart from the R200W mutation, 2 additional homozygous VHL mutations (Croatian H191D and P138L) and several (compound) heterozygous mutations have been discovered in single patients, resulting in very similar phenotypes observed in classical patients with Chuvash erythrocytosis.55-59  Conversely, the well-known autosomal dominant cancer predisposition to VHL, with >1500 known VHL mutations, does not lead to erythrocytosis and is caused by inheritance of a single mutated allele of VHL.60 

Since 2006, several patients and families with heterozygous loss-of-function mutations in the PHD2 gene have been described.55,61-64  The first mutations discovered were the P317R and the P371H variants, which affect the catalytic rate and substrate binding of PHD2, leading to partial inhibition of HIF hydroxylation.64-66  A few of the reported PHD2 mutations, apart from erythrocytosis, also led to other pathologic conditions such as superficial thrombophlebitis,64  sagittal sinus thrombosis,66  and hypertension.67  However, the number of such patients is currently still too small to draw firm conclusions. In only 1 case has PHD2 also been described to be associated with tumor formation—in particular to a recurrent paraganglioma. This patient with this tumor is a heterozygous carrier of a PHD2 germline mutation (H374R), which affects one of the 3 conserved amino acids that coordinate Fe2+ binding, therefore contributing to the functionality of the enzyme.68  Interestingly, sequence analysis of the removed tumor mass showed that not one but both PHD2 alleles were mutated in the tumor cells (loss of heterozygosity). Functional analysis of the described PHD2 variants revealed that only the H374R variant has a detrimental effect, and all other studied PHD2 mutations show only weak deficiency in HIFα regulation.63  Such functional differences may permit PHD2 to act as a tumor suppressor in patients.

A new form of familial erythrocytosis was discovered in a family where the phenotype was associated with a heterozygous missense mutation in the HIF2α gene (EPAS1). The mutation is predicted to produce a G537W change in the amino acid sequence of HIF2α, which is very close to the primary site of hydroxylation (Pro-531).69  The resulting impairment of the hydroxylation of HIF2α and its subsequent VHL binding leads to an aberrant stabilization of this transcription factor during normoxia. Further studies have revealed numerous other HIF2α alterations, all near the primary hydroxylation site, typically leading to elevated Epo levels and erythrocytosis in the affected patients.55,70  In addition, numerous single nucleotide polymorphisms in the HIF2α gene are found in Tibetans and are associated with only a moderate increase in hemoglobin concentrations. This adaptation to high altitude strengthens the link between HIF2α and erythropoiesis.71,72  Contrarily, mutations of the HIF1α isoform have not been associated with altered red blood cell production.

Interestingly, mutations of the HIF2α gene have not only been shown to lead to erythrocytosis but have also been recently described to cause neoplasia. In particular, 1 patient carrying an inherited gain-of-function mutation in HIF2α (F374Y) displayed erythrocytosis, with additional recurrent multiple paragangliomas.73  In addition, 2 erythrocytosis patients with paragangliomas, one who had an additional somatostatinoma, have also been described to carry somatic HIF2α mutations (A530T and A530V), which increase the half-life of the HIF2α subunit and enhance HIF downstream signaling.74  The mutation was found in DNA from the tumor cells only and not in other cell types nor in the patients’ parents, which argues for a causative postzygotic event.74  Screening of patients with chromaffin-cell tumors (i.e., paragangliomas, pheochromocytomas) led to the discovery of numerous other somatic HIF2α mutations that are only partially accompanied by erythrocytosis.75-78  This predicts a direct oncogenic role for HIF2α, independent of its impact on red blood cell production.

Taken together, patients bearing a polymorphism in VHL, PHD2, or HIF2α collectively highlight the importance of the HIF signaling pathway in red blood cell homeostasis. Both somatic and germline mutations in HIF pathway members have been shown to lead to erythrocytosis. In some cases, erythrocytosis was accompanied by neuroendocrine tumors whose molecular basis remains to be unraveled (Table 1).

Only a limited amount of erythrocytosis-associated mutations in the HIF pathway proteins in humans have been described thus far,most of them only very recently. To unravel the effective role of the different HIF pathway proteins during erythropoiesis, various genetically modified mice have been developed in the past 15 years (Table 2).

Table 2

Available genetically modified mice illustrating the impact of a deregulated HIF system on murine erythropoiesis

GeneType of modificationPhenotype miceRef.
VHL Liver-specific KO Erythrocytosis 95  
 Brain-specific KO Erythrocytosis 16  
 R200W mutation Erythrocytosis 98  
PHD2 Induced complete KO Erythrocytosis 88,89  
 Conditional KO (including Epo-producing cells in kidney and brain) Erythrocytosis 92  
 Heterozygosity Mild erythrocytosis 99  
HIF2α Complete KO Pancytopenia 100  
 Induced complete KO Anemia 82  
 Liver-specific KO Anemia 85 * 
83 * 
 Kidney-specific KO Anemia 83 
 Heterozygosity Mild anemia Unpublished 
 G536W mutation Erythrocytosis 86  
GeneType of modificationPhenotype miceRef.
VHL Liver-specific KO Erythrocytosis 95  
 Brain-specific KO Erythrocytosis 16  
 R200W mutation Erythrocytosis 98  
PHD2 Induced complete KO Erythrocytosis 88,89  
 Conditional KO (including Epo-producing cells in kidney and brain) Erythrocytosis 92  
 Heterozygosity Mild erythrocytosis 99  
HIF2α Complete KO Pancytopenia 100  
 Induced complete KO Anemia 82  
 Liver-specific KO Anemia 85 * 
83 * 
 Kidney-specific KO Anemia 83 
 Heterozygosity Mild anemia Unpublished 
 G536W mutation Erythrocytosis 86  

KO, knockout.

*

In anemic mice and during early postnatal development.

Although HIF1α was initially discovered as the isoform that activates EPO transcription,5  it was only after both systemic and cell type–specific HIF1α and HIF2α knockout mice were made that the distinct role of both these transcription factors in erythropoiesis became clear. HIF1α knockout mice (HIF1α−/−) are only viable up to E11.5, and these embryos show major defects of the cardiovascular system and the neural tube.79,80  However, the lack of HIF1α does not lead to complete abolishment of erythropoiesis, but rather to multiple disturbances in the adaptive responses to hypoxia. Conversely, HIF2α-deficient mice revealed that the observed pancytopenia is caused by abnormally low plasma Epo levels and impaired renal Epo induction.81  Ablation of this subunit after birth resulted in anemia accompanied by decreased circulatory Epo.82  Interestingly, even heterozygous-deficient mice (HIF2α+/−) show a mild form of anemia (unpublished data). The group of Haase and colleagues was able to demonstrate that the regulation of erythropoiesis is essentially driven by renal HIF2α.83  Indeed, specific deletion of HIF2α in the kidney resulted in Epo-dependent anemia, which was only partially compensated by hepatic HIF2α.83  Moreover, although both HIF isoforms are expressed in the kidney, only HIF2α is found in the peritubular interstitial cells,32,74  and co-localized with EPO mRNA in these cells.84  At the molecular level, it was shown that HIF2α is actually the major isoform binding the 3′ enhancer of the EPO gene in its native form, whereas HIF1α primarily binds to the isolated HRE, as initially described.4,5,85  Moreover, the existence of additional transcription factors that bind to sites outside the actual HRE, which promote the preferential binding of HIF2α, has been proposed as well.85  Recently, Lee and coworkers presented a new mouse line bearing a G536W missense mutation in HIF2α that corresponds to the first such human mutation identified (G537W). Remarkably, not only did these mice show elevated hematocrit and pulmonary hypertension, but these findings attest that missense mutations in HIF2α can indeed cause erythrocytosis.86 

The HIFα subunits are regulated by different PHDs—the oxygen sensors. However, it is only after the mutant mouse lines were created that the functional differences among the family members became clear. Systemic deletion of PHD2, leads to embryonic lethality caused by placental and heart defects, whereas PHD1 and PHD3-specific knockout mice do not show any apparent abnormalities.87  Inducible PHD2-deficient mice on the other hand, develop severe erythrocytosis and show decreased life expectancy.88,89  Mice that are systemically deficient for either PHD1 or PHD3 do not display increased hematocrit values, and only mice lacking both of these isoforms simultaneously develop a moderate form of erythrocytosis. In the latter mice, plasma EPO and renal EPO expression is decreased, whereas hepatic EPO mRNA is induced.89  Thus, PHD1 and PHD3 appear to have only minor roles in the regulation of EPO expression, although their additional loss in the background of PHD2 deficiency can ameliorate the erythrocytosis phenotype.90,91  Our research group recently developed a conditional PHD2-deficient mouse line displaying severe but nonlethal erythrocytosis.92  Using different genetic approaches (PHD2/HIFα double-deficient mice) we could show that the Epo-dependent red blood cell increase is driven by HIF2α, which is in line with other observations made in familial erythrocytosis.69,93  Conversely, we found that HIF1α actually serves as a protective factor in these PHD2-deficient mice via the local induction of PHD3.92 

Mice carrying a homozygous deletion of the VHL gene die in utero because of a defect in placental vasculogenesis.94  A liver-specific VHL deletion led to hepatic vascular tumors and erythrocytosis, which was accompanied by increased Epo levels.95  The increase in erythrocytes was not reversible by additional hepatic HIF1α deletion,96  but only by deletion of HIF2α.85  Mice with an astrocyte-specific deletion of VHL not only exhibited a significant increase in cerebral EPO mRNA, but also a significant induction of plasma Epo and erythrocytosis.16  The additional deletion of HIF1α did not correct this increase in red blood cell count but rather made the phenotype more severe and shortened the survival time of these double-deficient mice. On the other hand, elimination of HIF2α along with VHL normalized the red blood cell count and most of the cerebral EPO transcript.16  Recently, ablation of VHL in osteoblasts led to HIF2α-dependent EPO induction in these cells, accompanied by erythrocytosis and enhanced bone formation.97  In 2007, a mouse line carrying the homozygous R200W mutation (leading to Chuvash erythrocytosis in humans) was created. Interestingly, this point mutation resulted in moderate erythrocytosis accompanied by splenic erythropoiesis.98  Embryonic stem cells carrying this mutation exhibited normoxic stabilization of HIF2α, which was accompanied by upregulation of HIF2α targets like vascular endothelial growth factor.

Deregulation of EPO transcription caused by mutations in HIF pathway proteins is an important underlying cause of erythrocytosis in patients. Moreover, these mutations can also result in other pathologic conditions like tumor development. Recently, various point mutations in the HIF2α/EPAS1, VHL, and PHD2 genes have been identified, and additional studies have led to new insights into the HIF pathway. Complementary to these mutations, many genetically modified mice have provided a powerful tool to study the effect and location of HIF pathway members in relation to erythropoiesis and its additional risk factors. Furthermore, it might be of great interest to develop new mouse models for erythrocytosis and related diseases including mice that carry specific point mutations found in humans (as mentioned previously for the R200W VHL and very recently the G537W HIF2α mutation).

The authors thank the entire Wielockx laboratory for helpful discussions, and Dr. Soulafa Mamlouk and Dr. Vasuprada Iyengar in particular for critical reading of the manuscript.

This work is supported in part by grants from the MeDDrive-Programm (TU Dresden, Germany) and the Emmy Noether DFG (WI 3291/1-1, WI32911-2) (B.W.), and the Swiss National Science Foundation (M.G.).

This review is a collaborative work within the COST Action TD0901 “HypoxiaNet.” We apologize to our colleagues whose work was not cited because of space limitations.

Contribution: K.F. and B.W. wrote the manuscript; and M.G. provided helpful discussions and helped write the manuscript.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: Ben Wielockx, Emmy Noether Group (DFG) Institute of Pathology—University of Technology Dresden, Schubertstrasse 15, D-01307 Dresden, Germany; e-mail: Ben.Wielockx@uniklinikum-dresden.de.

1
Jourdanet
 
D
Influence de la Pression de L'air sur la Vie de L'homme
1875
Paris
Masson
2
Miyake
 
T
Kung
 
CK
Goldwasser
 
E
Purification of human erythropoietin.
J Biol Chem
1977
, vol. 
252
 
15
(pg. 
5558
-
5564
)
3
Jacobs
 
K
Shoemaker
 
C
Rudersdorf
 
R
, et al. 
Isolation and characterization of genomic and cDNA clones of human erythropoietin.
Nature
1985
, vol. 
313
 
6005
(pg. 
806
-
810
)
4
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
1992
, vol. 
12
 
12
(pg. 
5447
-
5454
)
5
Wang
 
GL
Jiang
 
BH
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
1995
, vol. 
92
 
12
(pg. 
5510
-
5514
)
6
Wang
 
GL
Semenza
 
GL
General involvement of hypoxia-inducible factor 1 in transcriptional response to hypoxia.
Proc Natl Acad Sci USA
1993
, vol. 
90
 
9
(pg. 
4304
-
4308
)
7
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
1994
, vol. 
91
 
14
(pg. 
6496
-
6500
)
8
Nagao
 
M
Ebert
 
BL
Ratcliffe
 
PJ
Pugh
 
CW
Drosophila melanogaster SL2 cells contain a hypoxically inducible DNA binding complex which recognises mammalian HIF-binding sites.
FEBS Lett
1996
, vol. 
387
 
2-3
(pg. 
161
-
166
)
9
Jacobson
 
LO
Goldwasser
 
E
Fried
 
W
Plzak
 
L
Role of the kidney in erythropoiesis.
Nature
1957
, vol. 
179
 
4560
(pg. 
633
-
634
)
10
Koury
 
ST
Bondurant
 
MC
Koury
 
MJ
Localization of erythropoietin synthesizing cells in murine kidneys by in situ hybridization.
Blood
1988
, vol. 
71
 
2
(pg. 
524
-
527
)
11
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 erythropoietin mRNA, and serum erythropoietin concentration.
Blood
1989
, vol. 
74
 
2
(pg. 
645
-
651
)
12
Bachmann
 
S
Le Hir
 
M
Eckardt
 
KU
Co-localization of erythropoietin mRNA and ecto-5′-nucleotidase immunoreactivity in peritubular cells of rat renal cortex indicates that fibroblasts produce erythropoietin.
J Histochem Cytochem
1993
, vol. 
41
 
3
(pg. 
335
-
341
)
13
Maxwell
 
PH
Osmond
 
MK
Pugh
 
CW
, et al. 
Identification of the renal erythropoietin-producing cells using transgenic mice.
Kidney Int
1993
, vol. 
44
 
5
(pg. 
1149
-
1162
)
14
Obara
 
N
Suzuki
 
N
Kim
 
K
Nagasawa
 
T
Imagawa
 
S
Yamamoto
 
M
Repression via the GATA box is essential for tissue-specific erythropoietin gene expression.
Blood
2008
, vol. 
111
 
10
(pg. 
5223
-
5232
)
15
Marti
 
HH
Wenger
 
RH
Rivas
 
LA
, et al. 
Erythropoietin gene expression in human, monkey and murine brain.
Eur J Neurosci
1996
, vol. 
8
 
4
(pg. 
666
-
676
)
16
Weidemann
 
A
Kerdiles
 
YM
Knaup
 
KX
, et al. 
The glial cell response is an essential component of hypoxia-induced erythropoiesis in mice.
J Clin Invest
2009
, vol. 
119
 
11
(pg. 
3373
-
3383
)
17
Sakanaka
 
M
Wen
 
TC
Matsuda
 
S
, et al. 
In vivo evidence that erythropoietin protects neurons from ischemic damage.
Proc Natl Acad Sci USA
1998
, vol. 
95
 
8
(pg. 
4635
-
4640
)
18
Soliz
 
J
Joseph
 
V
Soulage
 
C
, et al. 
Erythropoietin regulates hypoxic ventilation in mice by interacting with brainstem and carotid bodies.
J Physiol
2005
, vol. 
568
 
Pt 2
(pg. 
559
-
571
)
19
Miskowiak
 
K
Inkster
 
B
Selvaraj
 
S
Wise
 
R
Goodwin
 
GM
Harmer
 
CJ
Erythropoietin improves mood and modulates the cognitive and neural processing of emotion 3 days post administration.
Neuropsychopharmacology
2008
, vol. 
33
 
3
(pg. 
611
-
618
)
20
Richmond
 
TD
Chohan
 
M
Barber
 
DL
Turning cells red: signal transduction mediated by erythropoietin.
Trends Cell Biol
2005
, vol. 
15
 
3
(pg. 
146
-
155
)
21
Hattangadi
 
SM
Wong
 
P
Zhang
 
L
Flygare
 
J
Lodish
 
HF
From stem cell to red cell: regulation of erythropoiesis at multiple levels by multiple proteins, RNAs, and chromatin modifications.
Blood
2011
, vol. 
118
 
24
(pg. 
6258
-
6268
)
22
Bauer
 
A
Tronche
 
F
Wessely
 
O
, et al. 
The glucocorticoid receptor is required for stress erythropoiesis.
Genes Dev
1999
, vol. 
13
 
22
(pg. 
2996
-
3002
)
23
Broudy
 
VC
Lin
 
NL
Priestley
 
GV
Nocka
 
K
Wolf
 
NS
Interaction of stem cell factor and its receptor c-kit mediates lodgment and acute expansion of hematopoietic cells in the murine spleen.
Blood
1996
, vol. 
88
 
1
(pg. 
75
-
81
)
24
Lenox
 
LE
Perry
 
JM
Paulson
 
RF
BMP4 and Madh5 regulate the erythroid response to acute anemia.
Blood
2005
, vol. 
105
 
7
(pg. 
2741
-
2748
)
25
Ramos
 
P
Casu
 
C
Gardenghi
 
S
, et al. 
Macrophages support pathological erythropoiesis in polycythemia vera and β-thalassemia.
Nat Med
2013
, vol. 
19
 
4
(pg. 
437
-
445
)
26
Jewell
 
UR
Kvietikova
 
I
Scheid
 
A
Bauer
 
C
Wenger
 
RH
Gassmann
 
M
Induction of HIF-1alpha in response to hypoxia is instantaneous.
FASEB J
2001
, vol. 
15
 
7
(pg. 
1312
-
1314
)
27
Fandrey
 
J
Gorr
 
TA
Gassmann
 
M
Regulating cellular oxygen sensing by hydroxylation.
Cardiovasc Res
2006
, vol. 
71
 
4
(pg. 
642
-
651
)
28
Xia
 
X
Lemieux
 
ME
Li
 
W
, et al. 
Integrative analysis of HIF binding and transactivation reveals its role in maintaining histone methylation homeostasis.
Proc Natl Acad Sci USA
2009
, vol. 
106
 
11
(pg. 
4260
-
4265
)
29
Schödel
 
J
Oikonomopoulos
 
S
Ragoussis
 
J
Pugh
 
CW
Ratcliffe
 
PJ
Mole
 
DR
High-resolution genome-wide mapping of HIF-binding sites by ChIP-seq.
Blood
2011
, vol. 
117
 
23
(pg. 
e207
-
e217
)
30
Ivan
 
M
Kondo
 
K
Yang
 
H
, et al. 
HIFalpha targeted for VHL-mediated destruction by proline hydroxylation: implications for O2 sensing.
Science
2001
, vol. 
292
 
5516
(pg. 
464
-
468
)
31
Jaakkola
 
P
Mole
 
DR
Tian
 
YM
, et al. 
Targeting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation.
Science
2001
, vol. 
292
 
5516
(pg. 
468
-
472
)
32
Epstein
 
AC
Gleadle
 
JM
McNeill
 
LA
, et al. 
C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation.
Cell
2001
, vol. 
107
 
1
(pg. 
43
-
54
)
33
Berra
 
E
Benizri
 
E
Ginouvès
 
A
Volmat
 
V
Roux
 
D
Pouysségur
 
J
HIF prolyl-hydroxylase 2 is the key oxygen sensor setting low steady-state levels of HIF-1alpha in normoxia.
EMBO J
2003
, vol. 
22
 
16
(pg. 
4082
-
4090
)
34
Oehme
 
F
Ellinghaus
 
P
Kolkhof
 
P
, et al. 
Overexpression of PH-4, a novel putative proline 4-hydroxylase, modulates activity of hypoxia-inducible transcription factors.
Biochem Biophys Res Commun
2002
, vol. 
296
 
2
(pg. 
343
-
349
)
35
Wiesener
 
MS
Jürgensen
 
JS
Rosenberger
 
C
, et al. 
Widespread hypoxia-inducible expression of HIF-2alpha in distinct cell populations of different organs.
FASEB J
2003
, vol. 
17
 
2
(pg. 
271
-
273
)
36
Hu
 
C-J
Wang
 
L-Y
Chodosh
 
LA
Keith
 
B
Simon
 
MC
Differential roles of hypoxia-inducible factor 1alpha (HIF-1alpha) and HIF-2alpha in hypoxic gene regulation.
Mol Cell Biol
2003
, vol. 
23
 
24
(pg. 
9361
-
9374
)
37
van Patot
 
MC
Gassmann
 
M
Hypoxia: adapting to high altitude by mutating EPAS-1, the gene encoding HIF-2α.
High Alt Med Biol
2011
, vol. 
12
 
2
(pg. 
157
-
167
)
38
Branco-Price
 
C
Zhang
 
N
Schnelle
 
M
, et al. 
Endothelial cell HIF-1α and HIF-2α differentially regulate metastatic success.
Cancer Cell
2012
, vol. 
21
 
1
(pg. 
52
-
65
)
39
Raval
 
RR
Lau
 
KW
Tran
 
MGB
, et al. 
Contrasting properties of hypoxia-inducible factor 1 (HIF-1) and HIF-2 in von Hippel-Lindau-associated renal cell carcinoma.
Mol Cell Biol
2005
, vol. 
25
 
13
(pg. 
5675
-
5686
)
40
Petrella
 
BL
Lohi
 
J
Brinckerhoff
 
CE
Identification of membrane type-1 matrix metalloproteinase as a target of hypoxia-inducible factor-2 alpha in von Hippel-Lindau renal cell carcinoma.
Oncogene
2005
, vol. 
24
 
6
(pg. 
1043
-
1052
)
41
Covello
 
KL
Kehler
 
J
Yu
 
H
, et al. 
HIF-2alpha regulates Oct-4: effects of hypoxia on stem cell function, embryonic development, and tumor growth.
Genes Dev
2006
, vol. 
20
 
5
(pg. 
557
-
570
)
42
Chen
 
F
Zhang
 
W
Liang
 
Y
, et al. 
Transcriptome and network changes in climbers at extreme altitudes.
PLoS ONE
2012
, vol. 
7
 
2
pg. 
e31645
 
43
Sergeyeva
 
A
Gordeuk
 
VR
Tokarev
 
YN
Sokol
 
L
Prchal
 
JF
Prchal
 
JT
Congenital polycythemia in Chuvashia.
Blood
1997
, vol. 
89
 
6
(pg. 
2148
-
2154
)
44
Ang
 
SO
Chen
 
H
Gordeuk
 
VR
, et al. 
Endemic polycythemia in Russia: mutation in the VHL gene.
Blood Cells Mol Dis
2002
, vol. 
28
 
1
(pg. 
57
-
62
)
45
Ang
 
SO
Chen
 
H
Hirota
 
K
, et al. 
Disruption of oxygen homeostasis underlies congenital Chuvash polycythemia.
Nat Genet
2002
, vol. 
32
 
4
(pg. 
614
-
621
)
46
Russell
 
RC
Sufan
 
RI
Zhou
 
B
, et al. 
Loss of JAK2 regulation via a heterodimeric VHL-SOCS1 E3 ubiquitin ligase underlies Chuvash polycythemia.
Nat Med
2011
, vol. 
17
 
7
(pg. 
845
-
853
)
47
Bushuev
 
VI
Miasnikova
 
GY
Sergueeva
 
AI
, et al. 
Endothelin-1, vascular endothelial growth factor and systolic pulmonary artery pressure in patients with Chuvash polycythemia.
Haematologica
2006
, vol. 
91
 
6
(pg. 
744
-
749
)
48
Gordeuk
 
VR
Sergueeva
 
AI
Miasnikova
 
GY
, et al. 
Congenital disorder of oxygen sensing: association of the homozygous Chuvash polycythemia VHL mutation with thrombosis and vascular abnormalities but not tumors.
Blood
2004
, vol. 
103
 
10
(pg. 
3924
-
3932
)
49
Sable
 
CA
Aliyu
 
ZY
Dham
 
N
, et al. 
Pulmonary artery pressure and iron deficiency in patients with upregulation of hypoxia sensing due to homozygous VHL(R200W) mutation (Chuvash polycythemia).
Haematologica
2012
, vol. 
97
 
2
(pg. 
193
-
200
)
50
Gordeuk
 
VR
Prchal
 
JT
Vascular complications in Chuvash polycythemia.
Semin Thromb Hemost
2006
, vol. 
32
 
3
(pg. 
289
-
294
)
51
Liu
 
X
Hao
 
L
Zhang
 
S
, et al. 
Genetic repression of mouse VEGF expression regulates coagulation cascade.
IUBMB Life
2010
, vol. 
62
 
11
(pg. 
819
-
824
)
52
Sergueeva
 
AI
Miasnikova
 
GY
Okhotin
 
DJ
, et al. 
Elevated homocysteine, glutathione and cysteinylglycine concentrations in patients homozygous for the Chuvash polycythemia VHL mutation.
Haematologica
2008
, vol. 
93
 
2
(pg. 
279
-
282
)
53
Perrotta
 
S
Nobili
 
B
Ferraro
 
M
, et al. 
Von Hippel-Lindau-dependent polycythemia is endemic on the island of Ischia: identification of a novel cluster.
Blood
2006
, vol. 
107
 
2
(pg. 
514
-
519
)
54
Liu
 
E
Percy
 
MJ
Amos
 
CI
, et al. 
The worldwide distribution of the VHL 598C>T mutation indicates a single founding event.
Blood
2004
, vol. 
103
 
5
(pg. 
1937
-
1940
)
55
Lee
 
FS
Percy
 
MJ
The HIF pathway and erythrocytosis.
Annu Rev Pathol
2011
, vol. 
6
 (pg. 
165
-
192
)
56
Bond
 
J
Gale
 
DP
Connor
 
T
, et al. 
Dysregulation of the HIF pathway due to VHL mutation causing severe erythrocytosis and pulmonary arterial hypertension.
Blood
2011
, vol. 
117
 
13
(pg. 
3699
-
3701
)
57
Lanikova
 
L
Lorenzo
 
F
Yang
 
C
Vankayalapati
 
H
Drachtman
 
R
Divoky
 
V
Prchal
 
JT
Novel homozygous VHL mutation in exon 2 is associated with congenital polycythemia but not with cancer.
Blood
2013
, vol. 
121
 
19
(pg. 
3918
-
3924
)
58
Tomasic
 
NL
Piterkova
 
L
Huff
 
C
, et al. 
The phenotype of polycythemia due to Croatian homozygous VHL (571C>G:H191D) mutation is different from that of Chuvash polycythemia (VHL 598C>T:R200W).
Haematologica
2013
, vol. 
98
 
4
(pg. 
560
-
567
)
59
Pastore
 
Y
Jedlickova
 
K
Guan
 
Y
, et al. 
Mutations of von Hippel-Lindau tumor-suppressor gene and congenital polycythemia.
Am J Hum Genet
2003
, vol. 
73
 
2
(pg. 
412
-
419
)
60
Nordstrom-O’Brien
 
M
van der Luijt
 
RB
van Rooijen
 
E
, et al. 
Genetic analysis of von Hippel-Lindau disease.
Hum Mutat
2010
, vol. 
31
 
5
(pg. 
521
-
537
)
61
Albiero
 
E
Ruggeri
 
M
Fortuna
 
S
, et al. 
Analysis of the oxygen sensing pathway genes in familial chronic myeloproliferative neoplasms and identification of a novel EGLN1 germ-line mutation.
Br J Haematol
2011
, vol. 
153
 
3
(pg. 
405
-
408
)
62
Albiero
 
E
Ruggeri
 
M
Fortuna
 
S
, et al. 
Isolated erythrocytosis: study of 67 patients and identification of three novel germ-line mutations in the prolyl hydroxylase domain protein 2 (PHD2) gene.
Haematologica
2012
, vol. 
97
 
1
(pg. 
123
-
127
)
63
Ladroue
 
C
Hoogewijs
 
D
Gad
 
S
, et al. 
Distinct deregulation of the hypoxia inducible factor by PHD2 mutants identified in germline DNA of patients with polycythemia.
Haematologica
2012
, vol. 
97
 
1
(pg. 
9
-
14
)
64
Percy
 
MJ
Zhao
 
Q
Flores
 
A
, et al. 
A family with erythrocytosis establishes a role for prolyl hydroxylase domain protein 2 in oxygen homeostasis.
Proc Natl Acad Sci USA
2006
, vol. 
103
 
3
(pg. 
654
-
659
)
65
Pappalardi
 
MB
Martin
 
JD
Jiang
 
Y
, et al. 
Biochemical characterization of human prolyl hydroxylase domain protein 2 variants associated with erythrocytosis.
Biochemistry
2008
, vol. 
47
 
43
(pg. 
11165
-
11167
)
66
Percy
 
MJ
Furlow
 
PW
Beer
 
PA
Lappin
 
TRJ
McMullin
 
MF
Lee
 
FS
A novel erythrocytosis-associated PHD2 mutation suggests the location of a HIF binding groove.
Blood
2007
, vol. 
110
 
6
(pg. 
2193
-
2196
)
67
Al-Sheikh
 
M
Moradkhani
 
K
Lopez
 
M
Wajcman
 
H
Préhu
 
C
Disturbance in the HIF-1alpha pathway associated with erythrocytosis: further evidences brought by frameshift and nonsense mutations in the prolyl hydroxylase domain protein 2 (PHD2) gene.
Blood Cells Mol Dis
2008
, vol. 
40
 
2
(pg. 
160
-
165
)
68
Ladroue
 
C
Carcenac
 
R
Leporrier
 
M
, et al. 
PHD2 mutation and congenital erythrocytosis with paraganglioma.
N Engl J Med
2008
, vol. 
359
 
25
(pg. 
2685
-
2692
)
69
Percy
 
MJ
Furlow
 
PW
Lucas
 
GS
, et al. 
A gain-of-function mutation in the HIF2A gene in familial erythrocytosis.
N Engl J Med
2008
, vol. 
358
 
2
(pg. 
162
-
168
)
70
Percy
 
MJ
Chung
 
YJ
Harrison
 
C
, et al. 
Two new mutations in the HIF2A gene associated with erythrocytosis.
Am J Hematol
2012
, vol. 
87
 
4
(pg. 
439
-
442
)
71
Beall
 
CM
Cavalleri
 
GL
Deng
 
L
, et al. 
Natural selection on EPAS1 (HIF2alpha) associated with low hemoglobin concentration in Tibetan highlanders.
Proc Natl Acad Sci USA
2010
, vol. 
107
 
25
(pg. 
11459
-
11464
)
72
Simonson
 
TS
Yang
 
Y
Huff
 
CD
, et al. 
Genetic evidence for high-altitude adaptation in Tibet.
Science
2010
, vol. 
329
 
5987
(pg. 
72
-
75
)
73
Lorenzo
 
FR
Yang
 
C
Ng Tang Fui
 
M
, et al. 
A novel EPAS1/HIF2A germline mutation in a congenital polycythemia with paraganglioma.
J Mol Med (Berl)
2013
, vol. 
91
 
4
(pg. 
507
-
512
)
74
Zhuang
 
Z
Yang
 
C
Lorenzo
 
F
, et al. 
Somatic HIF2A gain-of-function mutations in paraganglioma with polycythemia.
N Engl J Med
2012
, vol. 
367
 
10
(pg. 
922
-
930
)
75
Yang
 
C
Sun
 
MG
Matro
 
J
, et al. 
Novel HIF2A mutations disrupt oxygen sensing, leading to polycythemia, paragangliomas, and somatostatinomas.
Blood
2013
, vol. 
121
 
13
(pg. 
2563
-
2566
)
76
Comino-Méndez
 
I
de Cubas
 
AA
Bernal
 
C
, et al. 
Tumoral EPAS1 (HIF2A) mutations explain sporadic pheochromocytoma and paraganglioma in the absence of erythrocytosis.
Hum Mol Genet
2013
, vol. 
22
 
11
(pg. 
2169
-
2176
)
77
Taïeb
 
D
Yang
 
C
Delenne
 
B
, et al. 
First Report of Bilateral Pheochromocytoma in the Clinical Spectrum of HIF2A-Related Polycythemia-Paraganglioma Syndrome.
J Clin Endocrinol Metab
2013
, vol. 
98
 
5
(pg. 
E908
-
E913
)
78
Toledo
 
RA
Qin
 
Y
Srikantan
 
S
, et al. 
In vivo and in vitro oncogenic effects of HIF2A mutations in pheochromocytomas and paragangliomas.
Endocr Relat Cancer
2013
, vol. 
20
 
3
(pg. 
349
-
359
)
79
Iyer
 
NV
Kotch
 
LE
Agani
 
F
, et al. 
Cellular and developmental control of O2 homeostasis by hypoxia-inducible factor 1 alpha.
Genes Dev
1998
, vol. 
12
 
2
(pg. 
149
-
162
)
80
Ryan
 
HE
Lo
 
J
Johnson
 
RS
HIF-1 alpha is required for solid tumor formation and embryonic vascularization.
EMBO J
1998
, vol. 
17
 
11
(pg. 
3005
-
3015
)
81
Scortegagna
 
M
Ding
 
K
Zhang
 
Q
, et al. 
HIF-2alpha regulates murine hematopoietic development in an erythropoietin-dependent manner.
Blood
2005
, vol. 
105
 
8
(pg. 
3133
-
3140
)
82
Gruber
 
M
Hu
 
C-J
Johnson
 
RS
Brown
 
EJ
Keith
 
B
Simon
 
MC
Acute postnatal ablation of Hif-2alpha results in anemia.
Proc Natl Acad Sci USA
2007
, vol. 
104
 
7
(pg. 
2301
-
2306
)
83
Kapitsinou
 
PP
Liu
 
Q
Unger
 
TL
, et al. 
Hepatic HIF-2 regulates erythropoietic responses to hypoxia in renal anemia.
Blood
2010
, vol. 
116
 
16
(pg. 
3039
-
3048
)
84
Paliege
 
A
Rosenberger
 
C
Bondke
 
A
, et al. 
Hypoxia-inducible factor-2alpha-expressing interstitial fibroblasts are the only renal cells that express erythropoietin under hypoxia-inducible factor stabilization.
Kidney Int
2010
, vol. 
77
 
4
(pg. 
312
-
318
)
85
Rankin
 
EB
Biju
 
MP
Liu
 
Q
, et al. 
Hypoxia-inducible factor-2 (HIF-2) regulates hepatic erythropoietin in vivo.
J Clin Invest
2007
, vol. 
117
 
4
(pg. 
1068
-
1077
)
86
Tan
 
Q
Kerestes
 
H
Percy
 
MJ
, et al. 
Erythrocytosis and Pulmonary Hypertension in a Mouse Model of Human HIF2A Gain-of-Function Mutation [published online ahead of print May 2, 2013].
J Biol Chem
 
doi:10.1182/jbc.M112.444059
87
Takeda
 
K
Ho
 
VC
Takeda
 
H
Duan
 
L-J
Nagy
 
A
Fong
 
G-H
Placental but not heart defects are associated with elevated hypoxia-inducible factor alpha levels in mice lacking prolyl hydroxylase domain protein 2.
Mol Cell Biol
2006
, vol. 
26
 
22
(pg. 
8336
-
8346
)
88
Minamishima
 
YA
Moslehi
 
J
Bardeesy
 
N
Cullen
 
D
Bronson
 
RT
Kaelin
 
WG
Somatic inactivation of the PHD2 prolyl hydroxylase causes polycythemia and congestive heart failure.
Blood
2008
, vol. 
111
 
6
(pg. 
3236
-
3244
)
89
Takeda
 
K
Aguila
 
HL
Parikh
 
NS
, et al. 
Regulation of adult erythropoiesis by prolyl hydroxylase domain proteins.
Blood
2008
, vol. 
111
 
6
(pg. 
3229
-
3235
)
90
Minamishima
 
YA
Moslehi
 
J
Padera
 
RF
Bronson
 
RT
Liao
 
R
Kaelin
 
WG
A feedback loop involving the Phd3 prolyl hydroxylase tunes the mammalian hypoxic response in vivo.
Mol Cell Biol
2009
, vol. 
29
 
21
(pg. 
5729
-
5741
)
91
Minamishima
 
YA
Kaelin
 
WG
Reactivation of hepatic EPO synthesis in mice after PHD loss.
Science
2010
, vol. 
329
 
5990
pg. 
407
 
92
Franke
 
K
Kalucka
 
J
Mamlouk
 
S
, et al. 
HIF-1α is a protective factor in conditional PHD2-deficient mice suffering from severe HIF-2α-induced excessive erythropoiesis.
Blood
2013
, vol. 
121
 
8
(pg. 
1436
-
1445
)
93
Percy
 
MJ
Beer
 
PA
Campbell
 
G
, et al. 
Novel exon 12 mutations in the HIF2A gene associated with erythrocytosis.
Blood
2008
, vol. 
111
 
11
(pg. 
5400
-
5402
)
94
Gnarra
 
JR
Ward
 
JM
Porter
 
FD
, et al. 
Defective placental vasculogenesis causes embryonic lethality in VHL-deficient mice.
Proc Natl Acad Sci USA
1997
, vol. 
94
 
17
(pg. 
9102
-
9107
)
95
Haase
 
VH
Glickman
 
JN
Socolovsky
 
M
Jaenisch
 
R
Vascular tumors in livers with targeted inactivation of the von Hippel-Lindau tumor suppressor.
Proc Natl Acad Sci USA
2001
, vol. 
98
 
4
(pg. 
1583
-
1588
)
96
Rankin
 
EB
Higgins
 
DF
Walisser
 
JA
Johnson
 
RS
Bradfield
 
CA
Haase
 
VH
Inactivation of the arylhydrocarbon receptor nuclear translocator (Arnt) suppresses von Hippel-Lindau disease-associated vascular tumors in mice.
Mol Cell Biol
2005
, vol. 
25
 
8
(pg. 
3163
-
3172
)
97
Rankin
 
EB
Wu
 
C
Khatri
 
R
, et al. 
The HIF signaling pathway in osteoblasts directly modulates erythropoiesis through the production of EPO.
Cell
2012
, vol. 
149
 
1
(pg. 
63
-
74
)
98
Hickey
 
MM
Lam
 
JC
Bezman
 
NA
Rathmell
 
WK
Simon
 
MC
von Hippel-Lindau mutation in mice recapitulates Chuvash polycythemia via hypoxia-inducible factor-2alpha signaling and splenic erythropoiesis.
J Clin Invest
2007
, vol. 
117
 
12
(pg. 
3879
-
3889
)
99
Mazzone
 
M
Dettori
 
D
Leite de Oliveira
 
R
, et al. 
Heterozygous deficiency of PHD2 restores tumor oxygenation and inhibits metastasis via endothelial normalization.
Cell
2009
, vol. 
136
 
5
(pg. 
839
-
851
)
100
Scortegagna
 
M
Ding
 
K
Oktay
 
Y
, et al. 
Multiple organ pathology, metabolic abnormalities and impaired homeostasis of reactive oxygen species in Epas1-/- mice.
Nat Genet
2003
, vol. 
35
 
4
(pg. 
331
-
340
)
Sign in via your Institution