When the Brain Yearns for Oxygen

 

Tristan Leua    Vera Schützholda    Joachim Fandreya    Katja B. Ferenza,b

 

aInstitute of Physiology, University of Duisburg-Essen, Essen, Germany, bCeNIDE (Center for Nanointegration Duisburg-Essen) University of Duisburg-Essen, Duisburg, Germany

 

 

 

 

Key Words

Oxygen sensing • Hypoxia-inducible factor (HIF) • Neuroscience • Nervous system

 

Abstract

Nearly 30 years ago hypoxia-inducible factor (HIF) was described as a protein complex bound to regulatory DNA sequences termed hypoxia response elements because HIF binding induced transcription of the erythropoietin gene under hypoxia. However, it soon became clear that HIF is part of a ubiquitous cellular oxygen sensing system, which ensures finely tuned control of HIF abundance and activity in dependence of the cellular oxygen tension. For their discoveries of how cells sense and adapt to oxygen availability Gregg L. Semenza, William G. Kaelin Jr. and Sir Peter J. Ratcliffe received the Nobel Prize in Physiology or Medicine 2019. The Nobel laureates’ pioneering work on cellular oxygen sensing has unraveled that HIF has numerous target genes reflecting its multiple functions in cellular metabolism and adaptation to different levels of oxygen. Importantly, HIF is also crucial for the development of the nervous system. HIF has an influence on different neural cell types regarding neurogenesis, maturation and apoptosis. Furthermore, HIF is involved in pathophysiological processes of the brain like stroke and Alzheimer’s disease resulting in the development of HIF-related therapeutic approaches.

 

 

Introduction

 

Cellular adaptation to hypoxia depends on the transcriptional activity of hypoxia-inducible factors (HIFs). HIFs regulate the expression of numerous target genes, which enable cells to survive at low oxygen levels. These genes are all part of the strategies of an organism to overcome hypoxia by inducing angiogenesis, glycolysis or erythropoiesis [1, 2]. Hypoxia affects cells under different circumstances: When the organism is exposed to high altitude at low atmospheric pO2, when cells are located at or around the core of tumors or in highly inflamed tissues, when cells reside in malperfused ischemic tissues or when cells face lumina of organs with low luminal oxygen tension like in the intestine or bladder. These tissues can display pO2 values of <15 mmHg [3-5] or even lower during inflammation when oxygen-consuming immune cells invade and tissue perfusion becomes compromised by edema formation. Furthermore, the inflammatory response itself, in particular cytokines and bacterial lipopolysaccharides can directly activate HIFs, e.g. via NF‑κB [6-8]. In the brain, HIF accumulation often correlates with neurodegenerative diseases like stroke or Alzheimer’s disease [9, 10].

Regulation of oxygen-dependent HIF accumulation and activity was unraveled backwards from the first HIF target gene erythropoietin (EPO), which controls red blood cell production depending on the tissue demand of oxygen: anemia and exposure to high altitude both increase EPO production while breathing pure oxygen can suppress its de-novo synthesis [11]. Induction of EPO gene expression by hypoxia primed the geneticist Gregg L. Semenza from John-Hopkins-University to dissect the regulatory elements of the EPO gene [12]. Epo has also always been of interest for nephrologists like Sir Peter J. Ratcliffe from Oxford University, because kidney failure leads to anemia due to impaired EPO production in the kidneys [13]. Finally, it was known that loss-of-function mutations in the tumor suppressor von Hippel-Lindau protein (pVHL) caused too much EPO production and polycythemia, which raised the interest of the oncologist William G. Kaelin Jr. from Harvard University in the regulation of HIF.

The three pioneers in HIF research Gregg L. Semenza, William G. Kaelin Jr. and Sir Peter J. Ratcliffe have been awarded with the Nobel Prize in Physiology or Medicine 2019 “for discoveries of how cells sense and adapt to oxygen availability” [The Nobel Foundation]. In addition to solving a basic physiological question, namely how cells sense oxygen, the award underlines the importance of the HIF system for the clinical exploitation and the human welfare. In this mini-review we will recapitulate the discovery, the function and regulation of HIFs and will extend their role in the development of the nervous system.

 

 

The Nobel laureates

 

The puzzle of cellular oxygen sensing was solved with the parallel efforts of three independent groups led by William G. Kaelin Jr., Sir Peter J. Ratcliffe and Gregg L. Semenza contemporaneously working on two different continents; in the beginning often in a head-to-head race but especially later more and more in dialog apparent in many co-authored manuscripts. Gradually they described all important key proteins and key enzymes that participate in the “oxygen sensing pathway”. Importantly, only the integration of their findings enabled them to depict the complex map of how every single cell senses oxygen and adapts its metabolism to the actual situation.

It all started with the detection of the first hypoxia response element (HRE) on DNA 3’ downstream from the EPO gene in hepatoma cells by Semenza meticulously dissecting the regulatory DNA of the EPO gene [12]. Ratcliffe immediately recognized that these HREs were not solely responsible for the control of the EPO gene but also many other genes and recognized HIF as part of a widespread oxygen sensing mechanism [14-16]. Subsequently, Semenza and Wang were able to describe HIF as the transcription factor complex bound to the HRE to enhance EPO synthesis [17]. Ratcliffe then demonstrated the wide dissemination among different cell types first for HRE [18] and later for HIF-1 and defined many novel HIF target genes in addition to EPO assigned to glycolytic pathways [19]. Another milestone was the first characterization of the HIF-DNA binding complex by the Semenza lab in 1995 as a basic-helix-loop-helix (bHLH) per-ARNT-Sim (PAS) heterodimer of a α- and β-subunit [20]. Afterwards, Ratcliffe and others perceived that the stability of HIF-α (and thus of the HIF-DNA binding complex) is regulated by oxygen-dependent poly-ubiquitination within the oxygen-dependent degradation domain (ODD) resulting in proteasomal degradation of this subunit [21, 22]. It was Ratcliffe again who, in 1999, identified E3 ubiquitin ligase as the operating enzyme with pVHL as the subunit crucial for binding HIF-α to the enzyme as a prerequisite for subsequent poly-ubiquitination [23], although the binding of the tumor suppressor pVHL to elongins (later identified as subunits of E3 ubiquitin ligase) had already been described in 1995 by Kaelin [24].

Kaelin looked on the system from a clinical perspective. He was the first to establish the clinical link between pVHL deficiency and an increased risk for tumor formation as he realized that in pVHL-deficient cells HIF-α is not degraded despite the presence of oxygen thus mimicking hypoxia [25]. However, it took until 2001 until the Ratcliffe and the Kaelin group independently described that hydroxylation of HIF-α by oxygen-dependent dioxygenases (prolyl-4-hydroxylases) is mandatory for binding pVHL [26, 27].

With that final link, the first general map (surely non-complete) of oxygen-dependent gene regulation was successfully pictured.

 

 

Regulation of HIF

 

HIFs are heterodimeric transcription factors composed of an α- and a β-subunit [28]. The β‑subunit, also known as aryl hydrocarbon receptor nuclear translocator (ARNT), is constitutively expressed and permanently present inside the nucleus of the cell. In contrast, the stabilization of the α‑subunit is oxygen-dependent. HIF-α has three known isoforms: HIF‑1α, HIF-2α and HIF-3α [29, 30]. HIF-1α and HIF-2α both translocate to the nucleus, form heterodimers with ARNT and are transcriptionally active under hypoxia [28, 31], while the function of HIF-3α is not resolved yet. The expression patterns of HIF-1α and -2α differ: HIF‑1α is ubiquitously expressed among all organs, while HIF-2α expression is limited to endothelium, kidney, pancreas, liver, heart, lungs and intestine as well as the brain [30, 32, 33].

 

Protein structure of HIFs

The protein structure of all isoforms of HIF-α and ARNT is similar: With their bHLH domain they bind HREs on the DNA and act as a transcription factor regulating the expression of specific target genes [20, 34, 35]. The PAS domain, named after similarity with Period (a drosophila gene), ARNT and Single-minded (also from drosophila), enables the subunits to interact with each other [20]. Only the α-subunit has an ODD, which contains two proline residues that can be hydroxylated for HIF-α regulation [22]. HIF-α has two transactivation domains: an N-terminal (N-TAD) and a C-terminal (C-TAD) one, respectively [36]. The C‑TAD is responsible for the binding of co-activators like p300 and the CREB-like binding protein (CBP) to HIF to recruit additional co-activators. Binding of these co-activators is mandatory for the expression of C-TAD dependent target genes. Hydroxylation at an asparagine residue in the C-TAD inhibits the interaction with these co-activators [37].

 

HIF regulation by prolyl hydroxylases

Once the prolyl hydroxylases (PHD), also called egg-laying-defective nine (EGLN) or HIF prolyl hydroxylases (HPH), have hydroxylated the conserved proline residues in the ODD (HIF-1α: Pro402 and Pro564, HIF-2α: Pro405 and Pro531 in humans), the HIF-α subunit is marked for proteasomal degradation [38]. In human and mouse three PHD isoforms 1 – 3 have been described [2]. Under normoxia, PHDs are enzymatically active using the co-substrates α-ketoglutarate, Fe2+, ascorbate and most importantly molecular oxygen, which makes the PHDs oxygen-dependent. Hydroxylated proline residues within the HIF-α proteins are recognized by pVHL. pVHL is part of a protein complex that includes elongin B, elongin C and cullin-2 and possesses ubiquitin ligase E3 activity. Upon pVHL binding to proline-hydroxylated HIF-α-subunits, they are poly-ubiquitinylated and sent to degradation by the proteasome [23]. In contrast, under hypoxia enzymatic activity of PHDs stops due to the lack of oxygen, HIF-α accumulates and translocates into the nucleus. Here, HIF-α dimerizes with its constitutively nuclear partner ARNT to bind to HREs and induce the cellular answer to hypoxia by regulating target gene transcription. Of note, among the HIF-target genes are also PHD2 and PHD3, thus forming a negative feedback loop in a way that HIF accumulation leads to increase PHD abundance and activity which then reduces further HIF‑α accumulation. This permits precise adjustment of HIF activity [39].

 

HIF regulation by factor-inhibiting HIF

Furthermore, HIF activity is inhibited by hydroxylation of an asparagine residue in the C-TAD domain (HIF-1α: Asn803, HIF-2α: Asn847 in humans) at very low oxygen concentrations. Like PHDs, this factor-inhibiting HIF (FIH or FIH-1; also HIF1AN) belongs to the oxygen-dependent dioxygenases using the same co-factors namely Fe2+, 2-oxoglutarate, and ascorbate for its activity [37, 40]. Hydroxylation of the C-terminal asparagine does not affect HIF-α stability but prevents the recruitment of the co-activators p300/CBP mandatory for the expression of C‑TAD-dependent genes [38]. The regulation of HIFs under normoxia and hypoxia is represented schematically in Fig. 1.

 

Fig. 1. If sufficient O2 is present (normoxia), the proline (Pro) and asparagine (Asn) residues of hypoxia-inducible factor (HIF)-α subunits are hydroxylated by prolyl hydroxylases (PHDs) and factor-inhibiting HIF (FIH). This leads to recognition by von Hippel-Lindau protein (pVHL), polyubiquitination and subsequent proteasomal degradation; in parallel, hydroxylation of the C-terminus of HIF-α prevents the binding of co-activators for transcription. With lack of oxygen (hypoxia) the hydroxylases are reduced in activity, resulting in accumulation and nuclear translocation of HIF-α and subsequent dimerization with HIF-β in the nucleus. The HIF complex in the nucleus recruits the co-activator p300/CBP and enables the transcription of HIF target genes.

 

 

Transcriptional activity of HIF

 

Transcriptional activity of HIFs is essential for the survival of cells and organisms under low oxygen levels. HIFs stimulate the expression of genes integrated in erythropoiesis, angiogenesis and glycolysis. But also pathways not obviously connected to hypoxia are affected by HIFs, e.g. apoptosis, pH regulation, inflammation, oncogenesis, histone demethylation and others [41]. It is assumed that several hundred genes per cell type are directly targeted by HIFs. All HIF target genes hold a consensus HRE sequence, which is used to identify novel HIF regulated genes [42]. Of note, however, not all consensus HREs confer hypoxia inducibility to their target genes.

 

Erythropoetin as target gene of HIF

The first and most prominent target gene of HIF is EPO. EPO is a glycoprotein hormone and controls the expansion and differentiation of erythroid progenitor cells to form red blood cells; it thus ensures sufficient oxygen capacity of the blood to distribute oxygen throughout the human body [17]. In adults, the kidney mainly produces EPO in the interstitium between renal tubules where anemia or inspiratory hypoxia cause HIF accumulation in interstitial cells to turn on EPO gene expression. HIF-1 was first described when bound to the EPO enhancer under hypoxia and allowed its isolation and identification [13]. Later, it was shown that at least in mice Epo is a HIF-2α-regulated gene in renal interstitial and hepatic cells [43-45], although, it seems that in some cells both HIF-1 and HIF-2 control Epo expression [11].

EPO is also expressed in different areas of the brain. Here, HIF-1 induces EPO in the hypoxic retina of mice and thereby protects against light-induced retinal degeneration [46]. In the brain, EPO appears as an important regulator of neuronal progenitor cell differentiation. Signalling of the EPO-receptor with EPO as the ligand in neurons, astrocytes, oligodendrocytes and microglia is indispensable for brain development [47, 48]. While primal expression of EPO is mostly restricted to neurons and astrocytes, post-ischemic EPO can also be found in microglia and endothelial cells [49]. Furthermore, EPO promotes neurogenesis in the subventricular zone and supports migration of neuronal precursors into the ischemic cortex and striatum of neonatal rats [50]. EPO also induces hippocampal neurogenesis in in vitro models of neonatal stroke and prevents glutamate-induced neuronal death [51, 52]. Various studies have shown the neuroprotective role of EPO in ischemic injury and neurodegenerative diseases, suggesting a promising therapeutic applicability of EPO in neuronal damage [53-60]. In contrast, the application of EPO protein in the adult lateral ventricle of the mouse brain resulted in a decreased number of neural stem cells, because cells were forced to differentiate into neural precursor cells [61]. In addition, erythrocytosis develops as a side effect of neuroprotective use of EPO which might enhance viscosity of the blood and thus impair brain perfusion [62].

 

Vascular endothelial growth factor as HIF target gene

Another important and well-studied HIF target gene is vascular endothelial growth factor (VEGF) [63]. In general, VEGF stimulates angiogenesis, the formation of new blood vessels, to provide oxygen in undersupplied tissues where HIF is active. In the neuronal context, VEGF promotes neurogenesis in vitro and in vivo [64, 65]. It is presumed that VEGF provides a mechanistic linkage between angiogenesis and neurogenesis to promote this neuroprotective effect [66]. VEGF preconditioning led to stem cell remodelling and attenuated the age-related decay of adult hippocampal neurons in mice [67]. Because VEGF is a HIF-regulated gene, conditional Hif-1α- or Hif-2α-knockout in neurons was expected to reduce Vegf expression in the brain [68]. However, Vegf is apparently expressed primarily in astrocytes under hypoxic conditions [69], and thus this finding of unchanged Vegf expression was not surprising in the case of a neuronal loss of HIF-1 or HIF-2.

 

Metabolic target genes of HIF

Finally, HIF adapts cellular metabolism to hypoxia by switching from oxidative to glycolytic metabolism. Genes that are involved in glucose transport or glycolysis are induced by HIF, for example glucose transporter-1 (GLUT1), pyruvate dehydrogenase kinase 1 (PDK1) and lactate dehydrogenase A (LDHA) [19, 70-72], shuttling pyruvate away from mitochondrial oxidation [73]. As the effects of HIF activation remind of the Warburg effect and VEGF induction stimulates tumor vascularization, HIF activity is a crucial factor in the development and progression of cancers [74]. In common human cancer types, an increased HIF-1α expression was found compared to normal tissue, including breast, colon, lung, ovary, prostate, uterus and brain [75].

 

 

Relevance of HIF activity in the brain

 

As outlined above, HIF plays a pivotal role in the development of the central nervous system. In general, the role of HIF obviously is not only functional but also of morphological relevance. This is reflected by for example the finding that the absence of both HIF-1α and HIF-2α leads to substantial defects in development particularly in the vascular system, which are in part connected to effects on the development of the sympathetic nervous system. Loss of HIF-1α in the cardiac outflow tract, right ventricle and atrium, pharyngeal mesoderm, peripheral neurons and hindlimbs impaired the survival and proliferation of pre- and post-ganglionic neurons of the sympathetic system [76]. These defects resulted in hypoplasia of the sympathetic ganglion chain and decreased the number of chromaffin cells in the adrenal medulla indicating a critical role HIF-1α during development of the sympathetic nervous system. Mice with a loss of HIF-2α displayed pronounced bradycardia despite normal morphological development of the circulatory system [77]. Related to that, HIF-2α plays an important role at post-vasculogenesis stages and is required for the remodeling of the primary vascular network into a mature hierarchy pattern [78]. The organ of Zuckerkandl (OZ), the principle source of catecholamine production in mammalian embryos was shown to intensively express HIF-2α. Thus, HIF-2α-deficient mice contained substantially reduced catecholamine levels, which prevented proper cardiac function causing prenatal death [77].

 

Impact of HIFs on different neural cell types

Furthermore, the loss of HIF-1α led to the formation of a hydrocephalus and a concomitant reduction of the number of neurons in the cortex in mice [79]. This also resulted in a dysfunction of spatial memory and a reduced ability to consolidate memory. While little is known about the importance of HIF for the development of different neural cell types, it was demonstrated that the loss of HIF-1α in astrocytes has a positive effect on neurons by protecting them from hypoxia-induced cell death [80]. Deletion of HIF-1α in neurons, on the other hand, promoted hypoxia-induced cell death [80]. In oligodendrocytes, HIF reduced the maturation of oligodendrocyte precursor cells and induced hypomyelination [81].

In the case of a direct role of HIF-2 for cells of the nervous system, a protective effect on neural stem cells and promotion of neurogenesis were observed in zebrafish, as HIF-2 influences the production of apoptosis inhibitors [82]. Mouse studies also showed that the stem cell regulator octamer binding transcription factor 4 (OCT4) was induced by HIF-2 and further that a loss of HIF-2α in astrocytes led to malformations of the retina as the process of retinal angiogenesis is disturbed [83, 84]. Another study also described the influence of HIF-2 on angiogenesis: The HIF signalling pathway controls angiogenesis in the entire brain and especially the oxygen regulation of neurons affects the physiological formation of new blood vessels in postnatal mice [85].

 

Pathophysiological aspects of the brain related to HIFs

HIFs are also relevant under pathophysiological conditions in the brain. In a mouse model for stroke, a combined neuronal knockout of HIF-1α and HIF-2α led to a decreased infarct size in the early acute phase during the first 24 hours of reperfusion. In the subsequent phase after 72 hours of reperfusion, no differences compared to the infarct size of the control group were visible. The authors of the study interpreted their results with the finding that a reduced expression of anti-survival genes in the early acute phase led to the loss of both HIF proteins and to increased apoptotic cell death and reduced angiogenesis [68]. These findings imply that a particular focus should be laid on the time-regulated activation or inhibition of hypoxia-regulated cytoprotective factors when considering HIF-directed therapeutics for stroke therapy.

The complex roles and functions of HIF and its target genes makes a therapeutic approach obviously not an easy task. It is possible that lowering the HIF-2 level may cause other problems, such as increased formation of glial scars or altered cell migration [84, 86]. In squamous epithelial cells from head and neck tumors, it was already shown that activation of epidermal growth factor receptor is regulated via HIF-2α under hypoxic conditions. This leads to more aggressive tumor growth, which in turn can be associated with increased motility and migration of the tumor cells [87]. Furthermore, hypoxia-dependent p75 neurotrophin receptor stabilization in glioblastoma cells leads to upregulation of HIF-1α and HIF-2α and thus to an increased migration potential [88].

 

Regulation of PHDs as a therapeutic approach

Possible therapeutic approaches in neuronal diseases include the controlled inhibition of PHDs to increase HIF-2 accumulation and thus promote the neuroprotective effects of the HIF-2 pathway. The protection of cortical neurons by PHD inhibitors correlated with enhanced VEGF expression. This encourages the use of PHD inhibitors as a therapeutic approach for preventing cell death in conditions that are associated with metabolic stress in the central nervous system [89]. A possible target protein in this case would be PHD2, since it has already been shown that this PHD isoform in particular is a regulator of HIF-2 in the brain [90]. Contrary, a targeted destabilization of HIF-2, e.g. by activation of PHDs, would presumably increase the astrocyte fraction and provide the immigrating neurons with a supportive basis for integration into the existing cellular network. This is of crucial importance in neuronal diseases, such as in stroke, since studies with rats have already shown that after ischemic injury only about 20% of new neurons that emerged from precursor cells survive in the long term, whereby only 0.2% of the originally lost neurons could be replaced by these cells [91]. This opens, however, a therapeutic route to reduce HIF-2 signalling for potential therapeutic benefits. In addition, unfavourable environmental factors, like a lack of trophic conditions and insufficient connections to supporting cells such as astrocytes, are possible causes of this high mortality rate of new-born neurons. Thus, the survival rate of the immigrated neurons could be promoted by increasing the PHD2 level, which finally supports the regeneration of damaged brain tissue.

 

 

Therapeutic potential of the HIF pathway

 

Today, clinical relevance of the HIF pathway is obvious and non-controversial, at least for the use of therapeutically stabilizing HIF-α to increase EPO production when the kidney fails to provide sufficient amounts of the hormone such as in the anemia of chronic kidney disease. The need to increase EPO production to compensate for the anemia has led to the first clinical approval of the PHD inhibitor Roxadustat (FG-4592), developed by FibroGen and Astellas/AstraZeneca, in December 2018 in China [92] and in September 2019 in Japan [93]. The drug is expected to be filed in USA (FDA) and Europe (EMA) soon [94]. So far, Roxadustat is approved to treat patients with chronic kidney disease that depend on dialysis [92]. Meanwhile, studies confirmed non-inferiority of orally applied Roxadustat to parenteral epoetin alfa in long-term dialysis patients [95] as well as safety and efficacy in patients suffering from chronic kidney disease but not yet dependent on dialysis [96]. Roxadustat as well as three other structure-related PHD inhibitors (GSK1278863, Molidustat and Vadadustat) currently used in clinical trials all potently inhibit (with very similar IC50 values) PHD2, the most important PHD isoform in humans [97]. Given the widespread importance of HIF-dependent regulation, it is likely that one will observe the opening up to numerous indications in order to treat other diseases caused by HIF dysregulation such as tissue ischemia, cardiovascular illness, kidney injury and transplantation.

However, a HIF-dependent therapy of neurological diseases appears to be a rather intricate affair. Analysis of pharmacologically or genetically induced changes in HIF activity resulted in controversial outcomes. In one study a neuron-specific HIF-1α deletion resulted in increased tissue damage after stroke [98], whereas others showed a reduced ischemic damage in a similar model [99]. Nevertheless, in brain research, therapeutic approaches also focus on oral PHD inhibitors [100, 101]. Concerning this, the PHD inhibitor FG-4497 indicated promising potential to prevent neuronal damage and vascular leakage after stroke [102] and GSK360A decreased post-stroke brain injury as well as sensory, motor, and cognitive behavioural deficits in rats [101]. Furthermore, other neurodegenerative diseases associated with neuronal loss and vascular dysfunction, e.g. Alzheimer´s disease, might come into focus for a treatment with FG-4497 [89].

 

 

Disclosure Statement

 

The authors have no conflicts of interest to declare

 

 

References

 

1 Wenger RH: Cellular adaptation to hypoxia: O2-sensing protein hydroxylases, hypoxia-inducible transcription factors, and O2-regulated gene expression. FASEB J 2002;16:1151-1162.
https://doi.org/10.1096/fj.01-0944rev

 

2 Fandrey J, Gorr TA, Gassmann M: Regulating cellular oxygen sensing by hydroxylation. Cardiovasc Res 2006;71:642-651.
https://doi.org/10.1016/j.cardiores.2006.05.005

 

3 Karhausen J, Furuta GT, Tomaszewski JE, Johnson RS, Colgan SP, Haase VH: Epithelial hypoxia-inducible factor-1 is protective in murine experimental colitis. J Clin Invest 2004;114:1098-1106.
https://doi.org/10.1172/JCI200421086

 

4 Fluck K, Breves G, Fandrey J, Winning S: Hypoxia-inducible factor 1 in dendritic cells is crucial for the activation of protective regulatory T cells in murine colitis. Mucosal Immunol 2016;9:379-390.
https://doi.org/10.1038/mi.2015.67

 

5 Van Welden S, Selfridge AC, Hindryckx P: Intestinal hypoxia and hypoxia-induced signalling as therapeutic targets for IBD. Nat Rev Gastroenterol Hepatol 2017;14:596-611.
https://doi.org/10.1038/nrgastro.2017.101

 

6 Frede S, Stockmann C, Freitag P, Fandrey J: Bacterial lipopolysaccharide induces HIF-1 activation in human monocytes via p44/42 MAPK and NF-kappaB. Biochem J 2006;396:517-527.
https://doi.org/10.1042/BJ20051839

 

7 Bonello S, Zahringer C, BelAiba RS, Djordjevic T, Hess J, Michiels C, Kietzmann T, Gorlach A: Reactive oxygen species activate the HIF-1alpha promoter via a functional NFkappaB site. Arterioscler Thromb Vasc Biol 2007;27:755-761.
https://doi.org/10.1161/01.ATV.0000258979.92828.bc

 

8 van Uden P, Kenneth NS, Rocha S: Regulation of hypoxia-inducible factor-1alpha by NF-kappaB. Biochem J 2008;412:477-484.
https://doi.org/10.1042/BJ20080476

 

9 Shi H: Hypoxia inducible factor 1 as a therapeutic target in ischemic stroke. Curr Med Chem 2009;16:4593-4600.
https://doi.org/10.2174/092986709789760779

 

10 Ogunshola OO, Antoniou X: Contribution of hypoxia to Alzheimer's disease: is HIF-1alpha a mediator of neurodegeneration? Cell Mol Life Sci 2009;66:3555-3563.
https://doi.org/10.1007/s00018-009-0141-0

 

11 Fandrey J: Oxygen-dependent and tissue-specific regulation of erythropoietin gene expression. Am J Physiol Regul Integr Comp Physiol 2004;286:R977-988.
https://doi.org/10.1152/ajpregu.00577.2003

 

12 Semenza GL, Traystman MD, Gearhart JD, Antonarakis SE: Polycythemia in transgenic mice expressing the human erythropoietin gene. Proc Natl Acad Sci U S A 1989;86:2301-2305.
https://doi.org/10.1073/pnas.86.7.2301

 

13 Santoro A: Anemia in renal insufficiency. Rev Clin Exp Hematol 2002;Suppl 1:12-20.

 

14 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 1991;266:15563-15566.

 

15 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 U S A 1991;88:5680-5684.
https://doi.org/10.1073/pnas.88.13.5680

 

16 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 U S A 1991;88:10553-10557.
https://doi.org/10.1073/pnas.88.23.10553

 

17 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;12:5447-5454.
https://doi.org/10.1128/MCB.12.12.5447

 

18 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 U S A 1993;90:2423-2427.
https://doi.org/10.1073/pnas.90.6.2423

 

19 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 U S A 1994;91:6496-6500.
https://doi.org/10.1073/pnas.91.14.6496

 

20 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 U S A 1995;92:5510-5514.
https://doi.org/10.1073/pnas.92.12.5510

 

21 Pugh CW, O'Rourke JF, Nagao M, Gleadle JM, Ratcliffe PJ: Activation of hypoxia-inducible factor-1; definition of regulatory domains within the alpha subunit. J Biol Chem 1997;272:11205-11214.
https://doi.org/10.1074/jbc.272.17.11205

 

22 Huang LE, Gu J, Schau M, Bunn HF: Regulation of hypoxia-inducible factor 1alpha is mediated by an O2-dependent degradation domain via the ubiquitin-proteasome pathway. Proc Natl Acad Sci U S A 1998;95:7987-7992.
https://doi.org/10.1073/pnas.95.14.7987

 

23 Maxwell PH, Wiesener MS, Chang GW, Clifford SC, Vaux EC, Cockman ME, Wykoff CC, Pugh CW, Maher ER, Ratcliffe PJ: The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature 1999;399:271-275.
https://doi.org/10.1038/20459

 

24 Kibel A, Iliopoulos O, DeCaprio JA, Kaelin WG, Jr.: Binding of the von Hippel-Lindau tumor suppressor protein to Elongin B and C. Science 1995;269:1444-1446.
https://doi.org/10.1126/science.7660130

 

25 Iliopoulos O, Kibel A, Gray S, Kaelin WG, Jr.: Tumour suppression by the human von Hippel-Lindau gene product. Nat Med 1995;1:822-826.
https://doi.org/10.1038/nm0895-822

 

26 Ivan M, Kondo K, Yang H, Kim W, Valiando J, Ohh M, Salic A, Asara JM, Lane WS, Kaelin WG, Jr.: HIFalpha targeted for VHL-mediated destruction by proline hydroxylation: implications for O2 sensing. Science 2001;292:464-468.
https://doi.org/10.1126/science.1059817

 

27 Jaakkola P, Mole DR, Tian YM, Wilson MI, Gielbert J, Gaskell SJ, von Kriegsheim A, Hebestreit HF, Mukherji M, Schofield CJ, Maxwell PH, Pugh CW, Ratcliffe PJ: Targeting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science 2001;292:468-472.
https://doi.org/10.1126/science.1059796

 

28 Wang GL, Semenza GL: Purification and characterization of hypoxia-inducible factor 1. J Biol Chem 1995;270:1230-1237.
https://doi.org/10.1074/jbc.270.3.1230

 

29 Gu YZ, Moran SM, Hogenesch JB, Wartman L, Bradfield CA: Molecular characterization and chromosomal localization of a third alpha-class hypoxia inducible factor subunit, HIF3alpha. Gene Expr 1998;7:205-213.

 

30 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 1alpha regulates the VEGF expression and is potentially involved in lung and vascular development. Proc Natl Acad Sci U S A 1997;94:4273-4278.
https://doi.org/10.1073/pnas.94.9.4273

 

31 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. Studies in ARNT-deficient cells. J Biol Chem 1996;271:15117-15123.
https://doi.org/10.1074/jbc.271.25.15117

 

32 Tian H, McKnight SL, Russell DW: Endothelial PAS domain protein 1 (EPAS1), a transcription factor selectively expressed in endothelial cells. Genes Dev 1997;11:72-82.
https://doi.org/10.1101/gad.11.1.72

 

33 Wiesener MS, Jurgensen JS, Rosenberger C, Scholze CK, Horstrup JH, Warnecke C, Mandriota S, Bechmann I, Frei UA, Pugh CW, Ratcliffe PJ, Bachmann S, Maxwell PH, Eckardt KU: Widespread hypoxia-inducible expression of HIF-2alpha in distinct cell populations of different organs. FASEB J 2003;17:271-273.
https://doi.org/10.1096/fj.02-0445fje

 

34 Wang GL, Semenza GL: General involvement of hypoxia-inducible factor 1 in transcriptional response to hypoxia. Proc Natl Acad Sci U S A 1993;90:4304-4308.
https://doi.org/10.1073/pnas.90.9.4304

 

35 Dachs GU, Patterson AV, Firth JD, Ratcliffe PJ, Townsend KM, Stratford IJ, Harris AL: Targeting gene expression to hypoxic tumor cells. Nat Med 1997;3:515-520.
https://doi.org/10.1038/nm0597-515

 

36 Jiang BH, Zheng JZ, Leung SW, Roe R, Semenza GL: Transactivation and inhibitory domains of hypoxia-inducible factor 1alpha. Modulation of transcriptional activity by oxygen tension. J Biol Chem 1997;272:19253-19260.
https://doi.org/10.1074/jbc.272.31.19253

 

37 Lando D, Peet DJ, Whelan DA, Gorman JJ, Whitelaw ML: Asparagine hydroxylation of the HIF transactivation domain a hypoxic switch. Science 2002;295:858-861.
https://doi.org/10.1126/science.1068592

 

38 Kaelin WG: Proline hydroxylation and gene expression. Annu Rev Biochem 2005;74:115-128.
https://doi.org/10.1146/annurev.biochem.74.082803.133142

 

39 del Peso L, Castellanos MC, Temes E, Martin-Puig S, Cuevas Y, Olmos G, Landazuri MO: The von Hippel Lindau/hypoxia-inducible factor (HIF) pathway regulates the transcription of the HIF-proline hydroxylase genes in response to low oxygenJ Biol Chem 2003;278:48690-48695.
https://doi.org/10.1074/jbc.M308862200

 

40 Lando D, Peet DJ, Gorman JJ, Whelan DA, Whitelaw ML, Bruick RK: FIH-1 is an asparaginyl hydroxylase enzyme that regulates the transcriptional activity of hypoxia-inducible factor. Genes Dev 2002;16:1466-1471.
https://doi.org/10.1101/gad.991402

 

41 Fandrey J, Schodel J, Eckardt KU, Katschinski DM, Wenger RH: Now a Nobel gas: oxygen. Pflugers Arch 2019;471:1343-1358.
https://doi.org/10.1007/s00424-019-02334-8

 

42 Wenger RH, Stiehl DP, Camenisch G: Integration of oxygen signaling at the consensus HRE. Sci STKE 2005;2005:re12.
https://doi.org/10.1126/stke.3062005re12

 

43 Warnecke C, Zaborowska Z, Kurreck J, Erdmann VA, Frei U, Wiesener M, Eckardt KU: Differentiating the functional role of hypoxia-inducible factor (HIF)-1alpha and HIF-2alpha (EPAS-1) by the use of RNA interference: erythropoietin is a HIF-2alpha target gene in Hep3B and Kelly cells. FASEB J 2004;18:1462-1464.
https://doi.org/10.1096/fj.04-1640fje

 

44 Scortegagna M, Ding K, Zhang Q, Oktay Y, Bennett MJ, Bennett M, Shelton JM, Richardson JA, Moe O, Garcia JA: HIF-2alpha regulates murine hematopoietic development in an erythropoietin-dependent manner. Blood 2005;105:3133-3140.
https://doi.org/10.1182/blood-2004-05-1695

 

45 Rankin EB, Biju MP, Liu Q, Unger TL, Rha J, Johnson RS, Simon MC, Keith B, Haase VH: Hypoxia-inducible factor-2 (HIF-2) regulates hepatic erythropoietin in vivo. J Clin Invest 2007;117:1068-1077.
https://doi.org/10.1172/JCI30117

 

46 Grimm C, Wenzel A, Groszer M, Mayser H, Seeliger M, Samardzija M, Bauer C, Gassmann M, Reme CE: HIF-1-induced erythropoietin in the hypoxic retina protects against light-induced retinal degeneration. Nat Med 2002;8:718-724.
https://doi.org/10.1038/nm723

 

47 Juul S: Erythropoietin in the central nervous system, and its use to prevent hypoxic-ischemic brain damage. Acta Paediatr Suppl 2002;91:36-42.
https://doi.org/10.1111/j.1651-2227.2002.tb02904.x

 

48 Nagai A, Nakagawa E, Choi HB, Hatori K, Kobayashi S, Kim SU: Erythropoietin and erythropoietin receptors in human CNS neurons, astrocytes, microglia, and oligodendrocytes grown in culture. J Neuropathol Exp Neurol 2001;60:386-392.
https://doi.org/10.1093/jnen/60.4.386

 

49 Bernaudin M, Marti HH, Roussel S, Divoux D, Nouvelot A, MacKenzie ET, Petit E: A potential role for erythropoietin in focal permanent cerebral ischemia in mice. J Cereb Blood Flow Metab 1999;19:643-651.
https://doi.org/10.1097/00004647-199906000-00007

 

50 Iwai M, Cao G, Yin W, Stetler RA, Liu J, Chen J: Erythropoietin promotes neuronal replacement through revascularization and neurogenesis after neonatal hypoxia/ischemia in rats. Stroke 2007;38:2795-2803.
https://doi.org/10.1161/STROKEAHA.107.483008

 

51 Morishita E, Masuda S, Nagao M, Yasuda Y, Sasaki R: Erythropoietin receptor is expressed in rat hippocampal and cerebral cortical neurons, and erythropoietin prevents in vitro glutamate-induced neuronal death. Neuroscience 1997;76:105-116.
https://doi.org/10.1016/S0306-4522(96)00306-5

 

52 Osredkar D, Sall JW, Bickler PE, Ferriero DM: Erythropoietin promotes hippocampal neurogenesis in in vitro models of neonatal stroke. Neurobiol Dis 2010;38:259-265.
https://doi.org/10.1016/j.nbd.2010.01.015

 

53 Gui DM, Yang Y, Li X, Gao DW: Effect of erythropoietin on the expression of HIF-1 and iNOS in retina in chronic ocular hypertension rats. Int J Ophthalmol 2011;4:40-43.

 

54 Parra AL, Rodriguez JC: Nasal neuro EPO could be a reliable choice for neuroprotective stroke treatment. Cent Nerv Syst Agents Med Chem 2012;12:60-68.
https://doi.org/10.2174/187152412800229143

 

55 Zhang G, Lehmann HC, Bogdanova N, Gao T, Zhang J, Sheikh KA: Erythropoietin enhances nerve repair in anti-ganglioside antibody-mediated models of immune neuropathy. PLoS One 2011;6:e27067.
https://doi.org/10.1371/journal.pone.0027067

 

56 Zhang J, Wang Q, Xiang H, Xin Y, Chang M, Lu H: Neuroprotection with erythropoietin in preterm and/or low birth weight infants. J Clin Neurosci 2014;21:1283-1287.
https://doi.org/10.1016/j.jocn.2013.10.040

 

57 Chong ZZ, Shang YC, Mu Y, Cui S, Yao Q, Maiese K: Targeting erythropoietin for chronic neurodegenerative diseases. Expert Opin Ther Targets 2013;17:707-720.
https://doi.org/10.1517/14728222.2013.780599

 

58 Maiese K, Chong ZZ, Shang YC, Wang S: Erythropoietin: new directions for the nervous system. Int J Mol Sci 2012;13:11102-11129.
https://doi.org/10.3390/ijms130911102

 

59 Shang YC, Chong ZZ, Wang S, Maiese K: Prevention of beta-amyloid degeneration of microglia by erythropoietin depends on Wnt1, the PI 3-K/mTOR pathway, Bad, and Bcl-xL. Aging (Albany NY) 2012;4:187-201.
https://doi.org/10.18632/aging.100440

 

60 Noh MY, Cho KA, Kim H, Kim SM, Kim SH: Erythropoietin modulates the immune-inflammatory response of a SOD1(G93A) transgenic mouse model of amyotrophic lateral sclerosis (ALS). Neurosci Lett 2014;574:53-58.
https://doi.org/10.1016/j.neulet.2014.05.001

 

61 Shingo T, Sorokan ST, Shimazaki T, Weiss S: Erythropoietin regulates the in vitro and in vivo production of neuronal progenitors by mammalian forebrain neural stem cells. J Neurosci 2001;21:9733-9743.
https://doi.org/10.1523/JNEUROSCI.21-24-09733.2001

 

62 Rabie T, Marti HH: Brain protection by erythropoietin: a manifold task. Physiology (Bethesda) 2008;23:263-274.
https://doi.org/10.1152/physiol.00016.2008

 

63 Liu Y, Cox SR, Morita T, Kourembanas S: Hypoxia regulates vascular endothelial growth factor gene expression in endothelial cells. Identification of a 5' enhancer. Circ Res 1995;77:638-643.
https://doi.org/10.1161/01.RES.77.3.638

 

64 Jin K, Zhu Y, Sun Y, Mao XO, Xie L, Greenberg DA: Vascular endothelial growth factor (VEGF) stimulates neurogenesis in vitro and in vivo. Proc Natl Acad Sci U S A 2002;99:11946-11950.
https://doi.org/10.1073/pnas.182296499

 

65 Han W, Song X, He R, Li T, Cheng L, Xie L, Chen H, Jiang L: VEGF regulates hippocampal neurogenesis and reverses cognitive deficits in immature rats after status epilepticus through the VEGF R2 signaling pathway. Epilepsy Behav 2017;68:159-167.
https://doi.org/10.1016/j.yebeh.2016.12.007

 

66 Palmer TD, Willhoite AR, Gage FH: Vascular niche for adult hippocampal neurogenesis. J Comp Neurol 2000;425:479-494.
https://doi.org/10.1002/1096-9861(20001002)425:4<479::AID-CNE2>3.0.CO;2-3

 

67 Licht T, Rothe G, Kreisel T, Wolf B, Benny O, Rooney AG, Ffrench-Constant C, Enikolopov G, Keshet E: VEGF preconditioning leads to stem cell remodeling and attenuates age-related decay of adult hippocampal neurogenesis. Proc Natl Acad Sci U S A 2016;113:E7828-E7836.
https://doi.org/10.1073/pnas.1609592113

 

68 Barteczek P, Li L, Ernst AS, Bohler LI, Marti HH, Kunze R: Neuronal HIF-1alpha and HIF-2alpha deficiency improves neuronal survival and sensorimotor function in the early acute phase after ischemic stroke. J Cereb Blood Flow Metab 2017;37:291-306.
https://doi.org/10.1177/0271678X15624933

 

69 Sinor AD, Irvin SM, Cobbs CS, Chen J, Graham SH, Greenberg DA: Hypoxic induction of vascular endothelial growth factor (VEGF) protein in astroglial cultures. Brain Res 1998;812:289-291.
https://doi.org/10.1016/S0006-8993(98)00976-7

 

70 Ebert BL, Firth JD, Ratcliffe PJ: Hypoxia and mitochondrial inhibitors regulate expression of glucose transporter-1 via distinct Cis-acting sequences. J Biol Chem 1995;270:29083-29089.
https://doi.org/10.1074/jbc.270.49.29083

 

71 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 1996;313:809-814.
https://doi.org/10.1042/bj3130809

 

72 Gleadle JM, Ratcliffe PJ: Induction of hypoxia-inducible factor-1, erythropoietin, vascular endothelial growth factor, and glucose transporter-1 by hypoxia: evidence against a regulatory role for Src kinase. Blood 1997;89:503-509.
https://doi.org/10.1182/blood.V89.2.503

 

73 Semenza GL: Regulation of metabolism by hypoxia-inducible factor 1. Cold Spring Harb Symp Quant Biol 2011;76:347-353.
https://doi.org/10.1101/sqb.2011.76.010678

 

74 Semenza GL: HIF-1 and tumor progression: pathophysiology and therapeutics. Trends Mol Med 2002;8:S62-67.
https://doi.org/10.1016/S1471-4914(02)02317-1

 

75 Semenza GL: Hypoxia, clonal selection, and the role of HIF-1 in tumor progression. Criv Rev Biochem Mol Biol 2000;35:71-103.
https://doi.org/10.1080/10409230091169186

 

76 Bohuslavova R, Cerychova R, Papousek F, Olejnickova V, Bartos M, Gorlach A, Kolar F, Sedmera D, Semenza GL, Pavlinkova G: HIF-1alpha is required for development of the sympathetic nervous system. Proc Natl Acad Sci U S A 2019;116:13414-13423.
https://doi.org/10.1073/pnas.1903510116

 

77 Tian H, Hammer RE, Matsumoto AM, Russell DW, McKnight SL: The hypoxia-responsive transcription factor EPAS1 is essential for catecholamine homeostasis and protection against heart failure during embryonic development. Genes Dev 1998;12:3320-3324.
https://doi.org/10.1101/gad.12.21.3320

 

78 Peng J, Zhang L, Drysdale L, Fong GH: The transcription factor EPAS-1/hypoxia-inducible factor 2alpha plays an important role in vascular remodeling. Proc Natl Acad Sci U S A 2000;97:8386-8391.
https://doi.org/10.1073/pnas.140087397

 

79 Tomita S, Ueno M, Sakamoto M, Kitahama Y, Ueki M, Maekawa N, Sakamoto H, Gassmann M, Kageyama R, Ueda N, Gonzalez FJ, Takahama Y: Defective brain development in mice lacking the Hif-1alpha gene in neural cells. Mol Cell Biol 2003;23:6739-6749.
https://doi.org/10.1128/MCB.23.19.6739-6749.2003

 

80 Vangeison G, Carr D, Federoff HJ, Rempe DA: The good, the bad, and the cell type-specific roles of hypoxia inducible factor-1 alpha in neurons and astrocytes. J Neurosci 2008;28:1988-1993.
https://doi.org/10.1523/JNEUROSCI.5323-07.2008

 

81 Yuen TJ, Silbereis JC, Griveau A, Chang SM, Daneman R, Fancy SPJ, Zahed H, Maltepe E, Rowitch DH: Oligodendrocyte-encoded HIF function couples postnatal myelination and white matter angiogenesis. Cell 2014;158:383-396.
https://doi.org/10.1016/j.cell.2014.04.052

 

82 Ko CY, Tsai MY, Tseng WF, Cheng CH, Huang CR, Wu JS, Chung HY, Hsieh CS, Sun CK, Hwang SP, Yuh CH, Huang CJ, Pai TW, Tzou WS, Hu CH: Integration of CNS survival and differentiation by HIF2alpha. Cell Death Differ 2011;18:1757-1770.
https://doi.org/10.1038/cdd.2011.44

 

83 Covello KL, Kehler J, Yu H, Gordan JD, Arsham AM, Hu CJ, Labosky PA, Simon MC, Keith B: HIF-2alpha regulates Oct-4: effects of hypoxia on stem cell function, embryonic development, and tumor growth. Genes Dev 2006;20:557-570.
https://doi.org/10.1101/gad.1399906

 

84 Duan LJ, Takeda K, Fong GH: Hypoxia inducible factor-2alpha regulates the development of retinal astrocytic network by maintaining adequate supply of astrocyte progenitors. PLoS One 2014;9:e84736.
https://doi.org/10.1371/journal.pone.0084736

 

85 Nasyrov E, Nolan KA, Wenger RH, Marti HH, Kunze R: The neuronal oxygen-sensing pathway controls postnatal vascularization of the murine brain. FASEB J 2019;33:12812-12824.
https://doi.org/10.1096/fj.201901385RR

 

86 Nagao K, Oka K: HIF-2 directly activates CD82 gene expression in endothelial cells. Biochem Biophys Res Commun 2011;407:260-265.
https://doi.org/10.1016/j.bbrc.2011.03.017

 

87 Wang X, Schneider A: HIF-2alpha-mediated activation of the epidermal growth factor receptor potentiates head and neck cancer cell migration in response to hypoxia. Carcinogenesis 2010;31:1202-1210.
https://doi.org/10.1093/carcin/bgq078

 

88 Tong B, Pantazopoulou V, Johansson E, Pietras A: The p75 neurotrophin receptor enhances HIF-dependent signaling in glioma. Exp Cell Res 2018;371:122-129.
https://doi.org/10.1016/j.yexcr.2018.08.002

 

89 Niatsetskaya Z, Basso M, Speer RE, McConoughey SJ, Coppola G, Ma TC, Ratan RR: HIF prolyl hydroxylase inhibitors prevent neuronal death induced by mitochondrial toxins: therapeutic implications for Huntington's disease and Alzheimer's disease. Antioxid Redox Signal 2010;12:435-443.
https://doi.org/10.1089/ars.2009.2800

 

90 Li L, Saliba P, Reischl S, Marti HH, Kunze R: Neuronal deficiency of HIF prolyl 4-hydroxylase 2 in mice improves ischemic stroke recovery in an HIF dependent manner. Neurobiol Dis 2016;91:221-235.
https://doi.org/10.1016/j.nbd.2016.03.018

 

91 Arvidsson A, Collin T, Kirik D, Kokaia Z, Lindvall O: Neuronal replacement from endogenous precursors in the adult brain after stroke. Nat Med 2002;8:963-970.
https://doi.org/10.1038/nm747

 

92 Dhillon S: Roxadustat: First Global Approval. Drugs 2019;79:563-572.
https://doi.org/10.1007/s40265-019-01077-1

 

93 Astellas, FibroGen: Evrenzo® (roxadustat) Tablets Approved in Japan for the Treatment of Anemia Associated with Chronic Kidney Disease in Dialysis Patients. Tokyo and San Francisco, Astellas FibroGen, 2019. URL: https://www.astellas.com/us/news/4536.

 

94 Sciesielski LK, Kirschner KM: ExActa HIF prolyl hydroxylase inhibitors-The new lifestyle drug? Acta Physiol (Oxf) 2019;227:e13370.
https://doi.org/10.1111/apha.13370

 

95 Chen N, Hao C, Liu BC, Lin H, Wang C, Xing C, Liang X, Jiang G, Liu Z, Li X, Zuo L, Luo L, Wang J, Zhao MH, Liu Z, Cai GY, Hao L, Leong R, Wang C, Liu C, et al.: Roxadustat Treatment for Anemia in Patients Undergoing Long-Term Dialysis. N Engl J Med 2019;381:1011-1022.
https://doi.org/10.1056/NEJMoa1901713

 

96 Chen N, Hao C, Peng X, Lin H, Yin A, Hao L, Tao Y, Liang X, Liu Z, Xing C, Chen J, Luo L, Zuo L, Liao Y, Liu BC, Leong R, Wang C, Liu C, Neff T, Szczech L, et al.: Roxadustat for Anemia in Patients with Kidney Disease Not Receiving Dialysis. N Engl J Med 2019;381:1001-1010.
https://doi.org/10.1056/NEJMoa1813599

 

97 Yeh TL, Leissing TM, Abboud MI, Thinnes CC, Atasoylu O, Holt-Martyn JP, Zhang D, Tumber A, Lippl K, Lohans CT, Leung IKH, Morcrette H, Clifton IJ, Claridge TDW, Kawamura A, Flashman E, Lu X, Ratcliffe PJ, Chowdhury R, Pugh CW, et al.: Molecular and cellular mechanisms of HIF prolyl hydroxylase inhibitors in clinical trials. Chem Sci 2017;8:7651-7668.
https://doi.org/10.1039/C7SC02103H

 

98 Baranova O, Miranda LF, Pichiule P, Dragatsis I, Johnson RS, Chavez JC: Neuron-specific inactivation of the hypoxia inducible factor 1 alpha increases brain injury in a mouse model of transient focal cerebral ischemia. J Neurosci 2007;27:6320-6332.
https://doi.org/10.1523/JNEUROSCI.0449-07.2007

 

99 Helton R, Cui J, Scheel JR, Ellison JA, Ames C, Gibson C, Blouw B, Ouyang L, Dragatsis I, Zeitlin S, Johnson RS, Lipton SA, Barlow C: Brain-specific knock-out of hypoxia-inducible factor-1alpha reduces rather than increases hypoxic-ischemic damage. J Neurosci 2005;25:4099-4107.
https://doi.org/10.1523/JNEUROSCI.4555-04.2005

 

100 Harten SK, Ashcroft M, Maxwell PH: Prolyl hydroxylase domain inhibitors: a route to HIF activation and neuroprotection. Antioxid Redox Signal 2010;12:459-480.
https://doi.org/10.1089/ars.2009.2870

 

101 Zhou J, Li J, Rosenbaum DM, Zhuang J, Poon C, Qin P, Rivera K, Lepore J, Willette RN, Hu E, Barone FC: The prolyl 4-hydroxylase inhibitor GSK360A decreases post-stroke brain injury and sensory, motor, and cognitive behavioral deficits. PLoS One 2017;12:e0184049.
https://doi.org/10.1371/journal.pone.0184049

 

102 Reischl S, Li L, Walkinshaw G, Flippin LA, Marti HH, Kunze R: Inhibition of HIF prolyl-4-hydroxylases by FG-4497 reduces brain tissue injury and edema formation during ischemic stroke. PLoS One 2014;9:e84767.
https://doi.org/10.1371/journal.pone.0084767