Skip to main content
Howard Hughes Medical Institute Author Manuscripts logoLink to Howard Hughes Medical Institute Author Manuscripts
. Author manuscript; available in PMC: 2011 Jul 21.
Published in final edited form as: Nat Rev Cancer. 2008 Nov 6;8(12):967–975. doi: 10.1038/nrc2540

The impact of O2 availability on human cancer

Jessica A Bertout 1,2, Shetal A Patel 1,3, M Celeste Simon 1,3,4
PMCID: PMC3140692  NIHMSID: NIHMS311210  PMID: 18987634

Abstract

During the last century, it has been established that regions within solid tumors experience mild to severe oxygen deprivation, due to aberrant vascular function. These hypoxic regions are associated with altered cellular metabolism, as well as increased resistance to radiation and chemotherapy. As discussed in this Timeline, over the past decade, work from many laboratories has elucidated the mechanisms by which hypoxia-inducible factors (HIFs) modulate tumor cell metabolism, angiogenesis, growth, and metastasis. The central role played by intra-tumoral hypoxia and HTF in these processes has made them attractive therapeutic targets in the treatment of multiple human malignancies.


Oxygen (O2) is required for aerobic metabolism to maintain intracellular bioenergetics and serve as an electron acceptor in many organic and inorganic reactions. Hypoxia, defined as reduced O2 levels, occurs in a variety of pathological conditions, including stroke, tissue ischemia, inflammation, and the growth of solid tumors. The beginnings of hypoxia research in tumor biology can be traced back to observations made in the early 20th century by Otto Warburg who demonstrated that, unlike normal cells, tumor cells favor glycolysis, independent of cellular oxygenation levels. He postulated that tumor growth is caused by mitochondrial dysfunction in neoplastic cells, forcing them to generate energy through glycolysis (reviewed in 1). This hypothesis appears to be incorrect, but a number of other molecular mechanisms promoting “aerobic glycolysis” have been proposed including mutations and epigenetic changes in genes encoding tumor suppressors (e.g. p53), oncogene activation (e.g. c-Myc), and hypoxic adaptations {Denko, 2008 #6606; Gatenby, 2004 #6608; Deberardinis, 2008 #6609.

Ambient air is 21% O2 (150 mm Hg); however, most mammalian tissues exist at 2%-9% O2 (on average 40 mm Hg). “Hypoxia” is usually defined as ≤ 2% O2, while severe hypoxia or “anoxia” is defined as ≤ 0.02% O2. In the decades following Warburg’s observation, scientists sought to determine whether hypoxic or anoxic cells could be found in mammalian tumors and how these cells affected radiation therapy. Of great interest to radiation biologists and oncologists, the radioprotective effect of anoxia in normal tissues was demonstrated in the 1940s by Lacassagne and Evans et al., using whole body anoxia in newborn rodents (reviewed in {Gray, 1953 #6395}). If a subset of tumor cells did in fact exist in an environment deprived of O2, then they might be responsible for tumor recurrence after radiation. This realization is still fueling research today, over 60 years later.

Early demonstrations of tumor hypoxia

Before hypoxic cells could be visualized in tumors, their presence was inferred by some astute observations. In 1955, Thomlinson and Gray studied histology sections of human lung tumors and observed cells growing in “cords” running parallel to vascularized stroma 2. In large cords, they noted a necrotic core surrounded by a region of viable cells neighboring a capillary vessel. They proposed that necrosis was due to insufficient O2 and nutrient supply to the rapidly expanding tissue. Regardless of cord size, detectable bands of live cancer cells lying between the necrotic core and the surrounding stroma were consistently 170 microns in width, approximately the calculated distance of O2 diffusion (145 microns). They proposed that the edges of such necrotic cores harbor viable hypoxic tumor cells 2. Other human tumors, such as cervical and renal carcinomas exhibited similar histological characteristics, suggesting that this was not unique to the lung (reviewed in 3). The histological specimens also suggested that tumor cells are exposed to an O2 gradient ranging from efficient oxygenation near the stroma decreasing gradually to near anoxia bordering the necrotic regions. These varying O2 concentrations have since been shown to have significantly different effects on cellular processes 4.

Further evidence for the existence of viable hypoxic tumor cells that may influence tumor responses to radiation was offered by Powers and Tholmach. They irradiated lymphosarcomas in mice breathing either ambient air or hyperbaric O2 (three atmospheres of pressure) before transplanting them to a new mouse host. They observed decreased tumor cell survival if transplanted from mice breathing hyperbaric O2 as compared to 21%, suggesting that certain tumor cells were protected by decreased O2 conditions, but sensitized by a surplus of O2 delivered to the host. Moreover, tumors irradiated in dead mice (likely more hypoxic than viable animals breathing room air), were much less sensitive to radiation 5. Around the same time, Churchill-Davidson made similarly encouraging observations in cancer patients treated with radiation in combination with hyperbaric O2 6. While these results were promising and suggestive of the existence of hypoxic cells within tumors, they did not provide direct evidence and it was not until the latter decades of the 20th century that precise techniques for measuring O2 levels were developed.

Even today, scientists lack an optimal method of direct tissue O2 measurement that is non-invasive, precise, and quantitative; but this goal is clearly within reach (reviewed in 7). In the 1980’s, in vivo polarographic measurements with electrodes and ex vivo cryospectrophotometric measurements of oxyhemoglobin provided some insight into the regional oxygenation status of tumors (reviewed in 8). Nuclear magnetic resonance spectroscopy of 31P signals detecting tumor cell metabolism and sensitizer-adducts such as misonidazole 9,10 selectively binding hypoxic cells were also used in an attempt to identify hypoxic regions (reviewed in 11). However, none of these methods were ideal. While electrode measurements were certainly the most direct form of O2 measurement available, some human tumors were not accessible to electrodes and the electrode diameters were too large to allow precise measurements of cellular oxygenation status. Instead, they likely measured average O2 tensions in tumor sections, which would not accurately reflect dramatic variations that occur over very small distances 12. Cryospectrophotometric measurements readily determine the oxyhemoglobin saturation of individual erythrocytes in tumor vessels ex vivo, an indirect assessment of tumor oxygenation 8. 31P NMR spectroscopy was found to lack sensitivity and technical difficulties still limited the use of sensitizer adducts 11. Using several of these methods, studies in both the 1980s and early 1990s found decreased average O2 tensions in human tumor tissue as compared to normal tissue, as predicted by previous work. Substantial inter-tumor variability was observed however, indicating that predictions of tumor oxygenation would be difficult and underscoring the urgent need for more precise and direct O2 measurement techniques easily used in clinical settings 8. In the late 1980s, an additional complication was revealed by Chaplin et al., using Hoechst 33432. This DNA binding stain has a short distribution half life but remains bound to tumor cell DNA, thus labeling cells based on their proximity to the blood supply. Using this technique coupled with cell sorting and responses to treatment, Chaplin differentiated between diffusion limited, chronic hypoxia and acute hypoxia resulting from transient changes in blood flow through tumor vasculature of larger tumors. Cells adjacent to blood vessels may suddenly and intermittently become hypoxic as the vessel becomes momentarily occluded during treatment 13. These two types of hypoxia may have different effects on treatment responses, a possibility that must be considered in the selection of cancer therapy protocols 14.

The effects of hypoxia on cancer

As described by Otto Warburg in the 1920s, rapidly dividing tumor cells display increased glycolysis, even in the presence of oxygen. As a consequence, lactic acid concentrations are elevated, acidifying the environment. Initially, scientists realized that exposure to hypoxia slowed tumor growth in mice, although early experiments were confounded by hypoxia-induced generalized weight loss in experimental animals. In 1954, Barach and Bickerman repeated these experiments in hypothyroid and acclimatized mice and found similar results without generalized weight loss, confirming that O2 deprivation impaired tumor growth 15. In extreme cases of prolonged anoxia, cells would undergo necrosis, as noted in Thomlinson and Gray’s tumor sections (reviewed in 16).

The major focus of hypoxia research for most of the 20th century was its role in cancer treatment responses. As early as 1909, Schwarz and colleagues noted that normal mammalian cells irradiated under conditions of hypoxia or anoxia were less sensitive to radiation than those irradiated in the presence of O2 17. At that time, the changes were attributed to blood flow differences, and not to oxygen availability. In 1923, Petry’s studies on the effects of oxygen on vegetable seeds were the first to note a correlation between oxygen and radiosensitivity (reviewed in 18). In the 1950s, pioneering work by Gray and his colleagues (as well as Churchill-Davidson) established the existence of a hypoxic radioresistance effect in mammalian tumors and sought to quantify it. Gray used Ehrlich mouse ascites tumor cells irradiated in vitro and in mice under various O2 conditions. Similar to what had been shown in plants and insect tissues, Gray found that the ascites tumor cells showed a dependence of radiation sensitivity on O2 tension both in vitro and in vivo. Sensitivity was about three times greater under well-oxygenated conditions as compared to anoxic conditions 19,20. Subsequently, Churchill-Davidson published a cytological evaluation of damage in patient tumors after radiation. One half of the tumor was irradiated while patients breathed pure O2 at three atmospheres pressure, the other half while patients were breathing room air. Significantly more damage was observed in the half irradiated under high O2 conditions 21. In 1960, Dewey used human embryo liver cells to show that cells grown in vitro under anoxic conditions exhibited approximately two and a half fold greater colony forming ability after irradiation than cells grown in air 22. Hewitt and Wilson irradiated mouse leukemia cells in live mice or two minutes after death and observed an identical sensitivity ratio as observed by Dewey 23. Several years later, the experiments by Powers and Tholmach described above, further confirmed these findings 5. Numerous other studies followed evaluating hypoxic effects on transplanted or spontaneous tumor regression after irradiation in mice, confirming earlier experiments (reviewed in 24). Similarly, hypoxic cells were later found to be more resistant to many commonly used chemotherapeutic agents 25,26.

However, the 1970s and 1980s witnessed a shift in opinions concerning the importance of hypoxia in clinical responses to cancer treatment. Firstly, clinical responses to hyperbaric O2 treatment27 and then later to hypoxic cell radiosensitizers 28 were disappointing. Furthermore, hypoxic tumor cells appeared to reoxygenate in response to fractionated radiation treatment and therefore would have substantially less impact on tumor cell responses after fractionated therapy than after a single radiation dose29,30. Finally, scientists found that intrinsic cellular radiosensitivity alone at low doses could explain the differences in clinical responses observed 31. Many became skeptical that the radiosensitization of hypoxic tumor cells was going to be clinically beneficial.

The next decade witnessed a second shift in conventional wisdom. With the demonstration in the 1980s and 1990s that, like other mammalian tumors, human tumors also contained radioresistant hypoxic cells came the realization that hypoxic cells could be targeted directly with hypoxia-specific cytotoxins. The combination of radiation with a hypoxia-specific cytotoxin would thus have the potential to destroy the entire tumor cell population. Such cytotoxins thus became the focus of intense investigation in the early 1990s 32.

The question of the role of hypoxia in cancer treatment instigated many studies, some of which revealed a number of other critical effects of hypoxia on tumor progression as discussed in detail later, including angiogenesis 33, metastatic potential 34,35, DNA replication 35,36 and reduced protein synthesis 37. Finally, in the decade immediately preceding the discovery of the transcription factor Hypoxia Inducible Factor (HIF, see below), the expression of a number of genes was determined to be stimulated by O2 deprivation. The earliest to be identified were Glucose Regulated Proteins (GRPs) and O2 Regulated Proteins (ORPs)37,38. One ORP was later identified as heme oxygenase 39, while two others were revealed to be the same as two GRPs38,40. Other genes induced by hypoxia include those regulating hematopoesis and the vasculature (VEGF41, IL-1α42, endothelin-1 43, PDGF44, and erythropoietin [EPO] 45), glycolysis 46, metastasis (Cathepsin L 47), and DNA damage responses (Gadd45, Gadd153 48). How the cell sensed O2 deprivation and how gene transcription was regulated by hypoxia were completely open questions until the discovery of HIFs in the 1990s and further elucidation of HIF-α independent effects of hypoxia on cellular processes over the next decades (see Figure 3).

Figure 3.

Figure 3

Hypoxia regulates other critical pathways that impact tumor progression in a HIF-independent fashion. A schematic diagram of signaling pathways which regulate mRNA translation is shown. These include the mTOR pathway and the integrated stress response. Hypoxia influences AMPK and REDD1 activity, which act upstream to inhibit mTOR kinase activity, resulting in a decrease in cap-dependent translation. Hypoxia also inhibits the ER resident kinase PERK, which phosphorylates eIF2α, resulting in global protein synthesis inhibition. While HIFs contribute to mTOR regulation via the induction of REDD1 during chronic hypoxia, all of these pathways can be regulated by O2 deprivation in a HIF-independent manner. Although the mechanisms for mTOR and PERK regulation by changes in O2 remain unclear, they represent important additional therapeutic targets.

Identification of Hypoxia Inducible Factors (HIFs)

HIFs are heterodimeric transcription factors consisting of an α and a β subunit 49. Systematic characterization of the EPO gene by multiple groups led to the identification of a cis-acting “hypoxia response element” (HRE; 5′-TACGTGCT-3′) in the 3′-flanking region of this locus that confers O2 regulation of EPO expression 50-53. “HIF-1” was subsequently described by Gregg Semenza and Guang Wang in 1992 as a nuclear factor that is induced by hypoxia, binds to the EPO HRE, and promotes transcriptional activation of EPO in O2-starved cells 54,55. The HIF-1 binding site within the EPO HRE was used for its purification via DNA affinity chromatography and cloning of cDNAs encoding the HIF-1α and HIF-1β subunits in 1995 49. This demonstrated that both subunits are members of the basic-helix-loop-helix polypeptide family that contain a “PAS” domain, initially characterized in the Drosophila melanogaster Per and Sim proteins and mammalian aryl hydrocarbon nuclear translocator (ARNT). While HIF-1α was a novel protein, HIF-1β was in fact identical to ARNT. HIF-1α protein and HIF-1 DNA-binding activity were rapidly induced by hypoxia (1% O2), and quickly decayed upon reoxygenation of cells to ambient O2 levels (21%). Finally, HIF-1α protein levels were shown to increase exponentially as HeLa cells were exposed to decreasing O2 concentrations (6-0.5%) with a half maximal response between 1.5 and 2% O2 56.

The HIF-α/β dimer binds to HREs (5′-G/ACGTG-3′) associated with numerous transcriptional target genes critical to systemic hypoxia responses, such as angiogenesis and erythropoiesis, and cellular hypoxia responses involving metabolism, proliferation, motility, and autophagy (see Figure 1). In fact, hypoxia may be one of the most potent inducers of autophagy, where cellular components are degraded and recycled to restore ATP levels (see 57). HIF target genes influence development, physiology, and numerous diseases. In mammals, three genes have been shown to encode HIF-α subunits that appear to be similarly regulated by O2 availability. Of the three HIF-α proteins, HIF-1α and HIF-2α have been the most extensively characterized (see 58). In addition to being identified first, HIF-1α is expressed ubiquitously, whereas HIF-2α and HIF-3α are found in a subset of tissues. HIF-2α, initially identified as endothelial PAS domain protein 1 (EPAS1), also binds to HREs and upregulates gene expression but is restricted to vascular endothelium, liver parenchyma, lung type II pneumocytes, and kidney epithelial cells 59-61. In contrast, HIF-3α or inhibitory PAS domain protein (IPAS) acts as a dominant negative regulator of HIF-1α and HIF-2α mediated transcription and is found at high levels in the thymus, cerebellar Purkinje cells, and corneal epithelium of the eye 62,63. Unlike the HIF-α proteins, HIF-1β/ARNT is constitutively expressed and insensitive to changes in O2 levels. While, HIF-α mRNA levels change in response to hypoxia in select cell types, HIF regulation is primarily based on posttranslational modification and protein stability (See Figure 2).

Figure 1.

Figure 1

Genes activated by hypoxia-inducible factors (HIFs) involved in tumor progression. Genes encoding proteins involved in numerous aspects of tumor initiation, growth, and metastasis are transcriptionally activated by either encoding HIF-1α or HIF-2α. Examples include: inflammatory cell recruitment (SDF-1α, CXCR4), proliferation (cyclin-D2, IGF-2), survival (VEGF, erythropoietin), metabolism/mitochondrial function (glycolytic enzymes, PDK-1), extracellular matrix function (fibronectin-1, collagen type-5), motility (c-MET, SPF-1α), angiogenesis (VEGF, PDGF), and pH regulation (carbonic anhydrase-9).

Figure 2.

Figure 2

Regulation of HIF-α subunits by O2 availability and other intracellular metabolites. In oxygen replete cells, the HIF-prolyl hydroxylases (PHDs) are active, resulting in the hydroxylation of proline residues in HIF-α and their targeted degradation via the pVHL-proteosome pathway. “Factor-inhibiting HIF” (FIH) hydroxylation of an asparagine residue in the C-terminus of the HIF-α subunit, blocks p300 co-factor recruitment. This results in the inactivation of HIF-α subunit transcriptional activity. All of these processes are inhibited when O2 levels decrease or cells exhibit increased levels of reactive oxygen species (ROS) and other metabolites such as fumerate, succinate, and nitric oxide (NO). Here, the HIF-α subunits are stabilized, they recruit co-activators such as p300, and activate HIF target genes, as described in Figure 1.

It is important to note that while the HIFs are critical mediators of the transcriptional response to hypoxia, a number of HIF-independent pathways also respond to changes in O2 tension (Figure 3). These responses allow cells to acutely adapt to the energetic demands of decreased O2 availability by limiting energy consuming processes such as protein synthesis. Our understanding of the importance and regulation of these pathways is only beginning to emerge.

HIFs and cellular O2 sensing

How the HIF pathway “senses” changes in O2 concentration has been intensively studied since the cloning of the HIF-1α and HIF-1β subunits in 1995. While much has been elucidated concerning O2 sensation by metazoans, key questions remain such as how HIF-independent pathways (Figure 3) perceive changes in O2 availability. Huang et al. determined that HIF-α subunit stability is largely regulated via a region containing 200 amino acids referred to as the oxygen-dependent degradation domain (ODD) 64. Direct analysis of the HIF-1α protein revealed that it is ubiquitinated and degraded by the 26S proteasome in O2 replete cells 64-66. Peter Ratcliffe, Patrick Maxwell, and colleagues demonstrated in 1999 that renal carcinoma cells lacking the von Hippel-Lindau tumor suppressor protein (pVHL) constitutively express both HIF-1α and HTF-2α as well as multiple HIF target genes 67. Hypoxic regulation of HIF could be restored upon reintroduction of plasmids encoding the pVHL polypeptide. pVHL is the substrate recognition component of a ubiquitin-protein ligase complex that includes ElonginB, ElonginC, Cul2, and Rbx1 67-71 . William Kaelin, Nikola Pavletich, and coworkers determined that this complex is related to the Skp1-Cul1-F-box (SCF) family of ubiquitin-protein ligases that catalyze the transfer of ubiquitin from ubiquitin conjugating enzymes to specific lysine residues of their substrates 72.

Biochemical studies performed by the research groups of Ratcliffe, Kaelin, and Frank Lee indicated that HIF-α interaction with pVHL requires the O2-, 2-oxoglutarate-, and iron-dependent hydroxylation of proline residue 564 (within the human HIF-1α ODD) by an enzymatic activity reminiscent of procollagen prolyl hydroxylases 73-75. A family of three HIF-specific prolyl hydroxylases was subsequently identified based on characterizing candidate genes that encode 2-oxoglutarate-dependent dioxygenases as well as biochemical purification 76-78. The mammalian enzymes, referred to as prolyl hydroxylase domain (PHD) proteins, are orthologs of the Cainorhabditis elegans egl-9 gene 76. Like vhl-1 mutants, egl-9 deficient worms constitutively express HIF-1α. Hydroxylation and pVHL-mediated polyubiquitination of HIF-α is inhibited in O2-deprived cells, resulting in HIF-α accumulation, translocation to the nucleus, ARNT dimerization, and target gene activation (see Figure 2). The ability of HIF-α to interact with coactivators such as p300/CBP exclusively under hypoxic conditions is also regulated by hydroxylation, in this case that of a C-terminal asparagine residue by Factor inhibiting HIF-1 (FIH-1) 79-82. FIH-1 is also an iron- and 2-oxoglutarate-dependent dioxygenase that likely fine tunes the regulation of HIF by O2 availability.

The role of HIFs in cancer

Based on the analysis of human cancer biopsies and experimental animal models, it has become increasingly clear that HIFs play a critical role in cancer progression (see 83-85). Immunohistochemical techniques have demonstrated that HIF-1α is overexpressed in a broad spectrum of human malignancies 84. Furthermore, HIF-1α accumulation has been associated with poor patient survival in early stage cervical cancer 86, breast cancer 87,88, oligodendroglioma 89, ovarian cancer 90, endometrial cancer 91, and oropharyngeal squamous cell carcinoma 92. Significant associations between HIF-2α overexpression and increased patient mortality have been reported for other diseases, including non-small-cell lung cancer (NSCLC) 93, neuroblastoma 94, astrocytoma 95, and head and neck squamous cell carcinoma 96. Interestingly, HIF-1α overexpression was associated with decreased patient mortality for head and neck cancer 97 and NSCLC 98, suggesting that the two HIF-α subunits can have opposing effects on disease progression depending on tumor type. Along these lines, HIF-1α expression gradually decreases, whereas HIF-2α expression increases as renal carcinomas develop in patients with VHL disease, the cancer susceptibility syndrome associated with germline VHL mutations 99,100. An association between HIF-3α expression and patient prognosis has not been examined.

Numerous experiments involving subcutaneous injection of murine and human cancer cell lines into immunodeficient mice have supported the notion that HIFs figure prominently in tumor growth. The first experiment was performed by the Ratcliffe group using ARNT-deficient mouse hepatoma cells 101. Compared with tumors generated from wildtype cells, mutant tumors grew more slowly and exhibited decreased angiogenesis. Randall Johnson and coworkers demonstrated that immortalized and transformed Hif-1α−/− mouse embryo fibroblasts also produced smaller tumors in subcutaneous models 102. David Livingston, Andrew Kung, and colleagues disrupted HIF transcriptional activity by introducing C-terminal HIF-1α peptides into human breast and colon cancer cells, resulting in decreased tumor growth in nude mice 103. This peptide functions by inhibiting the ability of HIF-1α to interact with the CBP and p300 cofactors. Consistent with VHL patient sample analysis (see above), inhibiting HIF-1α in renal carcinoma cells results in increased xenograft tumor mass, while HIF-2α inhibition results in decreased tumor growth in nude recipients 99. Ectopic expression of HIF-3α in hepatoma cells decreases tumor growth and angiogenesis in a subcutaneous tumor model, consistent with the dominant negative function of HIF-3α 63. Finally, downregulating HIF-2α via RNA interference resulted in decreased subcutaneous neuroblastoma growth 94. The studies in renal carcinoma, which were the first to directly demonstrate that HIF-1α and HIF-2α could have opposing effects on tumor growth, spurred a new interest into understanding the molecular differences between the α subunits. These observations also demonstrated that the relative contributions of the α subunits to tumor growth versus suppression are likely to be tissue specific.

While the injection of human cancer cells into immunocompromised mice has aided our understanding of the role of HIFs in cancer, this approach exclusively evaluates the terminal stages of malignancy. It fails to investigate the processes of tumor initiation, progression, and metastasis over a prolonged period of disease. The use of conditional alleles of murine Hif-1α and Hif-2α in spontaneous mouse models of cancer should greatly advance the field. Recently, Liao et al. have demonstrated a pivotal role for HIF-1α in pulmonary metastasis using a transgenic model of breast cancer 104. Here, HIF-1α was conditionally ablated in the mammary epithelium of mice expressing the polyoma middle T oncoprotein in mammary glands. HIF-1α deficiency resulted in delayed tumor onset, reduced tumor growth and somewhat fewer tumor blood vessels. However, the number of metastases to the lung was significantly decreased. These results indicate that HIF-1α is unnecessary for breast cancer initiation or blood vessel recruitment, but plays an unexpected role in promoting metastatic potential. A role for HIF-1α in promoting breast cancer metastasis was also described by Hiraga et al. 105. Validating observations made in mouse models through expression profiling and immunohistochemistry of patient samples will also be an important tool to extend experimental findings to the clinic 106.

A variety of genetic alterations that activate oncogenes and inactivate tumor suppressor genes result in increased HIF-1α expression, such as VHL, PTEN, and ARF loss of function and ERBB2 and SRC gain of function (reviewed in 84). However, one of the first links between tumor hypoxia and a specific cancer genetic program was that of p53. Regions of hypoxia and necrosis are common features of solid tumors. In 1996 Amato Giaccia, Thomas Graeber, and coworkers demonstrated that loss of the p53 tumor suppressor protein reduced hypoxia induced cell death in these lesions 107. The authors proposed that O2 deprivation provides a selective pressure within tumors for the clonal expansion of rare cells acquiring p53 mutations. The molecular mechanism(s) carrying the hypoxic signal to p53 was unclear until An et al showed that p53 was actually stabilized as a result of physical association with HIF-1α 108. More recent studies have supported these initial observations, although some controversy concerning interactions between the HIF-1α and p53 pathways remains to this day.

HIFs also impact the c-Myc pathway: whereas HIF-1α opposes c-Myc 109, a potent regulator of proliferation and anabolic metabolism, HIF-2α promotes c-Myc activity 110. These observations describe crosstalk between responses to changes in O2 availability and a key transcription factor regulating cell growth 111. Because c-Myc is commonly dysregulated in cancer, this is also an important interaction to consider when developing anticancer therapies.

Therapeutic implications of tumor hypoxia

As described previously, early work on the role of O2 levels in cancer centered on the observation that poorly perfused tissues are more resistant to radiation 112. It was later that Gray made the direct connection between tissue oxygenation and tumor cell radioresistance, a principle known as the “O2 enhancement effect” 19. Oxygen was thought to act as a direct radiosensitizer, increasing damage to DNA through the formation of free radicals. Indeed, experiments conducted by Richard Hodgkiss and colleagues in the late 1980s demonstrated that the oxygen effect required O2 to be present at the exact time of radiation, leading to the “oxygen fixation” hypothesis (see textbox 2) 113. The relative contribution of cellular responses to hypoxia versus the direct effect of free radicals to hypoxic radioresistance is still being deciphered.

Box 2: The “oxygen fixation” hypothesis.

Irradiation of cells generates free radicals either in DNA or water (H2O) molecules. Free radicals in DNA (DNA•) can react with available O2 to generate a peroxy-radical (DNA – OO•), thus chemically modifying the DNA (‘oxygen fixation’). In the absence of O2, the DNA radical will be reduced, restoring the DNA to its original composition (DNA – H). In addition to the direct effects on DNA, free radicals produced from H2O (H2O•) can further damage nearby DNA.

Box 2: The “oxygen fixation” hypothesis

Early attempts to exploit the oxygen effect for therapeutic benefit focused on interventions to increase tumor oxygenation during radiation. Hyperbaric O2, red blood cell transfusions and erythropoietin administration were all attempted as methods to increase tumor oxygenation. However, these approaches did not gain widespread use due to conflicting reports of their efficacy in clinical trials 114. In the 1970s, nitroimidazole derivatives, which act as molecular mimics for O2, were tested in conjunction with radiation. Despite initial promise, clinical trials with nitroimidazoles demonstrated limited benefit in part due to dose-limiting toxicities 115. As an alternative to increasing tumor oxygenation, more recent strategies have attempted to take advantage of hypoxia to selectively kill tumor cells. Tirapazamine, described over 20 years ago, is a prototypic example of a hypoxia-activated prodrug 116. Tirapazamine is specifically reduced in hypoxic cells, forming radical species that poison topoisomerase II leading to DNA double-strand breaks 117. Clinical trials with this agent have demonstrated benefit in patients with lung as well as head and neck cancer, when used in combination with radiation or chemotherapy 118,119.

In addition to the direct role of molecular O2 in generating radiation induced DNA damage, the biological effects of hypoxia on tumor cells can also modulate their response to therapy. This was first demonstrated several decades ago in experiments showing that the duration of hypoxia influenced radiation sensitivity 120. Collectively, hypoxia modulates multiple pathways that contribute to radiation and chemotherapy resistance. Hypoxia exerts a selective pressure on cells for loss of p53, a key mediator of apoptosis, and upregulates the expression of the multidrug resistance gene (MDR1), leading to efflux of chemotherapeutic drugs 121-123. Hypoxia has also been shown to contribute to increased mutagenesis rates and metastasis 124-127. Lysyl oxidase, c-MET, and CXCR4 are examples of genes directly regulated by hypoxia that play important roles in metastasis, and could serve as druggable targets.

Given the impact of hypoxia on therapeutic efficacy and patient prognosis, the ability to measure and image hypoxia in tumors has garnered significant attention. The introduction of the polarographic O2 microelectrode (Eppendorf electrode) in the 1990s allowed the direct and most accurate measurement of O2 tension in tumors 128. However, due to the invasive nature of this technique and limitation to superficial tumors it cannot be broadly utilized in the clinic. Thus the emphasis has shifted to imaging modalities to indirectly observe tumor hypoxia. Increased glycolysis by tumors can be imaged using positron emission tomography (PET) with 18fluorodeoxyglucose (FDG) as a tracer. However, since many tumors display increased rates of aerobic glycolysis, this is not a selective marker for tumor hypoxia. Several other compounds have also been developed that selectively accumulate in hypoxic tissues, including fluoromisonidazole (FMISO), etanidazole penta-fluoride (EF-5), and Cu(II)-diacetyl-bis(N4-methylthiosemicarbazone (Cu-ATSM) 129-131. Nitroimidazoles are reduced under hypoxic conditions, forming stable adducts in the cell that can be detected by the PET scanner. While many of these approaches are currently being investigated for their ability to predict treatment response, they still have not attained routine use in the clinic.

The cloning of HIF, understanding of the pathways that regulate it, and identification of downstream HIF targets has produced numerous molecular targets for therapeutic intervention. Targeting the HIF pathway has generated significant attention because of the potential to selectively target hypoxic tumor cells and the role that hypoxia plays in key tumorigenic processes such as metastasis, angiogenesis and metabolism. Due to the difficulty in directly targeting transcription factors, the most promising interventions include strategies that block HIF accumulation, either by inhibiting translation or promoting degradation. Another approach is to inhibit key HIF target genes, such as VEGF, TGFα, or lysyl oxidase. Mutations in the PI3K/AKT/mTOR and RAS signaling pathways are commonly found in human tumors and both of these pathways have been shown to upregulate HIF-1α protein. Indeed, inhibition of the mTOR kinase with RAD-001 in a mouse model of prostate cancer induced by oncogenic Akt was found to decrease HIF-1α accumulation and HIF-dependent gene expression 132. Renal carcinoma cells, which often exhibit mutations in the VHL tumor suppressor, have also been shown to downregulate HIF-1α expression in response to mTOR inhibition 133. In addition to the hypoxia-regulated VHL-dependent HIF degradation pathway, HIF is degraded through interaction with receptor of activated protein kinase C (RACK-1) 134. RACK-1 competes with Hsp90 for binding to HIF-1α and Hsp90 inhibitors have been shown to promote HIF degradation. Targeting the interaction between HIF and the transcriptional co-activator p300 has also been demonstrated to inhibit HIF target gene expression and impair tumor growth 103,135. Numerous other small molecules, including topoisomerase inhibitors, microtubule destabilizers, and histone deacetylase inhibitors, inhibit HIF-1α accumulation, although the mechanism for these effects is not yet clear 136.

Finally, drugs against several key HIF transcriptional targets have been developed and approved for clinical use. Chief among these is bevacizumab (Avastin, Genentech), a monoclonal antibody against VEGF, a key endothelial growth factor involved in tumor angiogenesis 137. Bevacizumab is now commonly used for the treatment of metastatic colorectal cancer 138. VEGF activity can also be inhibited by VEGF receptor tyrosine kinase inhibitors such as sorafenib, which has shown clinical benefit in renal and hepatocellular carcinoma 139. The therapeutic effects of anti-VEGF therapy appear to involve several mechanisms, often depending on the tumor type and chemotherapy regimen. These effects include inhibition of angiogenesis, endothelial cell apoptosis, and vessel normalization 140. Additionally, epidermal growth factor receptor inhibitors such as Iressa have seen some success in the treatment of lung cancer141,142.

The past century of hypoxia research has provided significant advancement in our understanding of the molecular pathways by which O2 tension influences the properties of tumors, from proliferation and angiogenesis to radiation and chemotherapy resistance. Current studies are focused on extending observations made in vitro to more physiologic contexts. The cloning of the HIFs has provided a genetic tool that when coupled with established mouse models of cancer will help clarify HIF-dependent effects on disease progression. Accordingly, studies in which one or both of the HIF-α subunits is conditionally deleted are now underway and will allow us to determine the tissue and stage specific effects of these factors on tumor progression. Mouse models will also be a useful platform to test newly identified therapies that target the HIF pathway for the treatment of cancer. The recent identification of a number of hypoxia-regulated microRNAs provides another mechanism by which O2 levels can influence gene expression. A number of these microRNAs are highly expressed in tumors and their predicted targets (eg. BID, VEGF, cyclin D2) have roles in apoptosis, angiogenesis and proliferation. In future studies it will be important to identify the targets of these microRNAs to understand their role in tumorigenesis. 143. Nearly a century after Warburg’s initial observation of aerobic glycolysis in tumors, the unique metabolism of cancer cells will be important to reexamine. Specifically, the crosstalk between the HIF pathway and dysregulated oncogene and tumor suppressor function in driving metabolic changes will be an important area of study. Collectively, these observations will be critical for the successful development of novel imaging and treatment modalities, with the promise to improve targeted therapy of cancer and overcome therapeutic resistance induced by hypoxia.

Box 1: Timeline.

The history of research on tumor hypoxia

  • 1909 Schwartz first observed the effect of changes in vascular function on radiation sensitivity.

  • 1923 First correlation between oxygen and radiosensitivity made by Petry in Germany.

  • 1924 Otto Warburg describes increased tumor cell glycolysis in the presence of high O2 levels.

  • 1953 Gray and colleagues establish the existence of hypoxic radioresistance.

  • 1955 Thomlinson and Gray describe tumor cell hypoxia in clinical specimens.

  • 1964 Hyperbaric O2 used to promote the efficacy of radiation treatment of cancer patients.

  • 1984 First “oxygen regulated proteins” (ORPs) described.

  • 1986 Polarographic electrodes and nuclear magnetic resonance spectroscopy used to measure tumor cell oxygenation.

  • 1986 First attempts to exploit tumor cell hypoxia by selectively killing O2-starved cells with tirapazamine.

  • 1991 Hypoxia regulated DNA elements characterized in human erythropoietin (EPO) gene.

  • 1995 “Hypoxia Inducible Factor-1” (HIF-1) biochemically purified and cloned by Semenza and colleagues.

  • 1996 Giaccia and coworkers determine that tumor cell hypoxia selects for rare clones exhibiting p53 inactivating mutations.

  • 1997 First demonstration of tumor promoting effects by HIF in subcutaneous mouse models.

  • 1999 HIF regulation by the von Hippel-Lindau tumor suppressor protein (pVHL) described by Ratcliffe et al.

  • 2001 Ratcliffe, Kaelin, and Lee research groups demonstrate HIF-α subunit prolyl hydroxylation.

  • 2001 HIF-prolyl hydroxylase domain (PHD) enzymes identified.

Box 3: HIF contributions to other diseases.

  • -

    Polycythemia: Point mutations in HIF-2α have recently been identified in multiple families with erythrocytosis 144,145. A point mutation in VHL leading to Chuvash polycythemia preferentially stabilizes HIF-2α 146.

  • -

    Cardiovascular disease: Cardiac ischemia is a significant source of patient mortality. The adaptive responses mediated by the HIFs are important for protecting ischemic tissues from injury and activating the HIF pathway selectively in these tissues could provide an important therapeutic intervention 147-149.

  • -

    Pulmonary hypertension: Both HIF-1α and HIF-2α regulate vascular remodeling and development of pulmonary hypertension in response to chronic hypoxia 150,151.

  • -

    Inflammation: Hypoxia is commonly observed in inflammatory states, and HIF-1α can have protective as well as pro-inflammatory effects. Macrophage specific deletion of HIF-1α revealed an important role in mediating innate immune responses 152. Alternatively, deletion of HIF-1α in colonic epithelium revealed a protective function in drug-induced colitis 153.

References

  • 1.Warburg O. On the origin of cancer cells. Science. 1956;123:309–14. doi: 10.1126/science.123.3191.309. [DOI] [PubMed] [Google Scholar]
  • 2.Thomlinson RH, Gray LH. The histological structure of some human lung cancers and the possible implications for radiotherapy. Br J Cancer. 1955;9:539–49. doi: 10.1038/bjc.1955.55. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Thomlinson RH. Tumour anoxia and the response to radiation. Sci Basis Med Annu Rev. 1965:74–90. [PubMed] [Google Scholar]
  • 4.Simon MC, Liu L, Barnhart BC, Young RM. Hypoxia-induced signaling in the cardiovascular system. Annu Rev Physiol. 2008;70:51–71. doi: 10.1146/annurev.physiol.70.113006.100526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Powers WE, Tolmach LJ. Demonstration of an Anoxic Component in a Mouse Tumor-Cell Population by in Vivo Assay of Survival Following Irradiation. Radiology. 1964;83:328–36. doi: 10.1148/83.2.328. [DOI] [PubMed] [Google Scholar]
  • 6.Churchill-Davidson I. Oxygenation in Radiotherapy of Malignant Disease of the Upper Air Passages, the Oxygen Effect of Radiotherapy. Proc R Soc Med. 1964;57:635–8. [PMC free article] [PubMed] [Google Scholar]
  • 7.Okunieff P, Fenton B, Chen Y. Past, present, and future of oxygen in cancer research. Adv Exp Med Biol. 2005;566:213–22. doi: 10.1007/0-387-26206-7_29. [DOI] [PubMed] [Google Scholar]
  • 8.Vaupel P, Schlenger K, Hoeckel M. Blood flow and tissue oxygenation of human tumors: an update. Adv Exp Med Biol. 1992;317:139–51. doi: 10.1007/978-1-4615-3428-0_14. [DOI] [PubMed] [Google Scholar]
  • 9.Urtasun RC, Chapman JD, Raleigh JA, Franko AJ, Koch CJ. Binding of 3H-misonidazole to solid human tumors as a measure of tumor hypoxia. Int J Radiat Oncol Biol Phys. 1986;12:1263–7. doi: 10.1016/0360-3016(86)90273-7. [DOI] [PubMed] [Google Scholar]
  • 10.Urtasun RC, Koch CJ, Franko AJ, Raleigh JA, Chapman JD. A novel technique for measuring human tissue p02 at the cellular level. Br J Cancer. 1986;54:453–7. doi: 10.1038/bjc.1986.197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Chapman JD. Measurement of tumor hypoxia by invasive and non-invasive procedures: a review of recent clinical studies. Radiother Oncol. 1991;20(Suppl 1):13–9. doi: 10.1016/0167-8140(91)90181-f. [DOI] [PubMed] [Google Scholar]
  • 12.Helmlinger G, Yuan F, Dellian M, Jain RK. Interstitial pH and pO2 gradients in solid tumors in vivo: high-resolution measurements reveal a lack of correlation. Nat Med. 1997;3:177–182. doi: 10.1038/nm0297-177. [DOI] [PubMed] [Google Scholar]
  • 13.Chaplin DJ, Durand RE, Olive PL. Acute hypoxia in tumors: implications for modifiers of radiation effects. Int J Radiat Oncol Biol Phys. 1986;12:1279–82. doi: 10.1016/0360-3016(86)90153-7. [DOI] [PubMed] [Google Scholar]
  • 14.Dewhirst MW, Cao Y, Moeller B. Cycling hypoxia and free radicals regulate angiogenesis and radiotherapy response. Nat Rev Cancer. 2008;8:425–37. doi: 10.1038/nrc2397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Barach AL, Bickerman HA. The effect of anoxia on tumor growth with special reference to sarcoma 180 implanted in C57 mice. Cancer Res. 1954;14:672–6. [PubMed] [Google Scholar]
  • 16.Thomlinson RH. Hypoxia and tumours. J Clin Pathol Suppl (R Coll Pathol) 1977;11:105–13. doi: 10.1136/jcp.s3-11.1.105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Churchill-Davidson I, Sanger C, Thomlinson RH. High-pressure oxygen and radiotherapy. Lancet. 1955;268:1091–5. doi: 10.1016/s0140-6736(55)90589-4. [DOI] [PubMed] [Google Scholar]
  • 18.Hall EJ, Giaccia A. Radiobiology for the Radiologist. Lippincott Williams & Wilkins. 2006 [Google Scholar]
  • 19.Gray LH, Conger AD, Ebert M, Hornsey S, Scott OC. The concentration of oxygen dissolved in tissues at the time of irradiation as a factor in radiotherapy. Br J Radiol. 1953;26:638–48. doi: 10.1259/0007-1285-26-312-638. [DOI] [PubMed] [Google Scholar]
  • 20.Deschner EE, Gray LH. Influence of oxygen tension on x-ray-induced chromosomal damage in Ehrlich ascites tumor cells irradiated in vitro and in vivo. Radiat Res. 1959;11:115–46. [PubMed] [Google Scholar]
  • 21.Churchill-Davidson I, Sanger C, Thomlinson RH. Oxygenation in radiotherapy. II. Clinical application. Br J Radiol. 1957;30:406–22. doi: 10.1259/0007-1285-30-356-406. [DOI] [PubMed] [Google Scholar]
  • 22.Dewey DL. Effect of oxygen and nitric oxide on the radio-sensitivity of human cells in tissue culture. Nature. 1960;186:780–2. doi: 10.1038/186780a0. [DOI] [PubMed] [Google Scholar]
  • 23.Hewitt HB, Wilson CW. The effect of tissue oxygen tension on the radiosensitivity of leukaemia cells irradiated in situ in the livers of leukaemic mice. Br J Cancer. 1959;13:675–84. doi: 10.1038/bjc.1959.75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Gray LH. Radiobiologic basis of oxygen as a modifying factor in radiation therapy. Am J Roentgenol Radium Ther Nucl Med. 1961;85:803–15. [PubMed] [Google Scholar]
  • 25.Teicher BA. Hypoxia and drug resistance. Cancer Metastasis Rev. 1994;13:139–68. doi: 10.1007/BF00689633. [DOI] [PubMed] [Google Scholar]
  • 26.Teicher BA, Lazo JS, Sartorelli AC. Classification of antineoplastic agents by their selective toxicities toward oxygenated and hypoxic tumor cells. Cancer Res. 1981;41:73–81. [PubMed] [Google Scholar]
  • 27.Henk JM. Does hyperbaric oxygen have a future in radiation therapy? Int J Radiat Oncol Biol Phys. 1981;7:1125–8. doi: 10.1016/0360-3016(81)90173-5. [DOI] [PubMed] [Google Scholar]
  • 28.Brown JM. Clinical trials of radiosensitizers: what should we expect? Int J Radiat Oncol Biol Phys. 1984;10:425–9. doi: 10.1016/0360-3016(84)90063-4. [DOI] [PubMed] [Google Scholar]
  • 29.Kallman RF, Dorie MJ. Tumor oxygenation and reoxygenation during radiation therapy: their importance in predicting tumor response. Int J Radiat Oncol Biol Phys. 1986;12:681–5. doi: 10.1016/0360-3016(86)90080-5. [DOI] [PubMed] [Google Scholar]
  • 30.Coleman CN. Modulating the radiation response. Stem Cells. 1996;14:10–5. doi: 10.1002/stem.140010. [DOI] [PubMed] [Google Scholar]
  • 31.Steel GG, McMillan TJ, Peacock JH. The radiobiology of human cells and tissues. In vitro radiosensitivity. The picture has changed in the 1980s. Int J Radiat Biol. 1989;56:525–37. doi: 10.1080/09553008914551691. [DOI] [PubMed] [Google Scholar]
  • 32.Brown JM, Giaccia AJ. Tumour hypoxia: the picture has changed in the 1990s. Int J Radiat Biol. 1994;65:95–102. doi: 10.1080/09553009414550131. [DOI] [PubMed] [Google Scholar]
  • 33.Blumenson LE, Bross ID. A possible mechanism for enhancement of increased production of tumor angiogenic factor. Growth. 1976;40:205–9. [PubMed] [Google Scholar]
  • 34.van den Brenk HA, Moore V, Sharpington C, Orton C. Production of metastases by a primary tumour irradiated under aerobic and anaerobic conditions in vivo. Br J Cancer. 1972;26:402–12. doi: 10.1038/bjc.1972.53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Young SD, Marshall RS, Hill RP. Hypoxia induces DNA overreplication and enhances metastatic potential of murine tumor cells. Proc Natl Acad Sci U S A. 1988;85:9533–7. doi: 10.1073/pnas.85.24.9533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Young SD, Hill RP. Effects of reoxygenation on cells from hypoxic regions of solid tumors: analysis of transplanted murine tumors for evidence of DNA overreplication. Cancer Res. 1990;50:5031–8. [PubMed] [Google Scholar]
  • 37.Heacock CS, Sutherland RM. Enhanced synthesis of stress proteins caused by hypoxia and relation to altered cell growth and metabolism. Br J Cancer. 1990;62:217–25. doi: 10.1038/bjc.1990.264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Sciandra JJ, Subjeck JR, Hughes CS. Induction of glucose-regulated proteins during anaerobic exposure and of heat-shock proteins after reoxygenation. Proc Natl Acad Sci U S A. 1984;81:4843–7. doi: 10.1073/pnas.81.15.4843. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Murphy BJ, Laderoute KR, Short SM, Sutherland RM. The identification of heme oxygenase as a major hypoxic stress protein in Chinese hamster ovary cells. Br J Cancer. 1991;64:69–73. doi: 10.1038/bjc.1991.241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Roll DE, Murphy BJ, Laderoute KR, Sutherland RM, Smith HC. Oxygen regulated 80 kDa protein and glucose regulated 78kDa protein are identical. Mol Cell Biochem. 1991;103:141–8. doi: 10.1007/BF00227480. [DOI] [PubMed] [Google Scholar]
  • 41.Shweiki D, Itin A, Soffer D, Keshet E. Vascular endothelial growth factor induced by hypoxia may mediate hypoxia-initiated angiogenesis. Nature. 1992;359:843–845. doi: 10.1038/359843a0. [DOI] [PubMed] [Google Scholar]
  • 42.Shreeniwas R, et al. Hypoxia-mediated induction of endothelial cell interleukin-1 alpha. An autocrine mechanism promoting expression of leukocyte adhesion molecules on the vessel surface. J Clin Invest. 1992;90:2333–9. doi: 10.1172/JCI116122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Kourembanas S, Marsden PA, McQuillan LP, Faller DV. Hypoxia induces endothelin gene expression and secretion in cultured human endothelium. J Clin Invest. 1991;88:1054–7. doi: 10.1172/JCI115367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Kourembanas S, Hannan RL, Faller DV. Oxygen tension regulates the expression of the platelet-derived growth factor-B chain gene in human endothelial cells. J Clin Invest. 1990;86:670–4. doi: 10.1172/JCI114759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Goldberg MA, Dunning SP, Bunn HF. Regulation of the erythropoietin gene: evidence that the oxygen sensor is a heme protein. Science. 1988;242:1412–5. doi: 10.1126/science.2849206. [DOI] [PubMed] [Google Scholar]
  • 46.Robin ED, Murphy BJ, Theodore J. Coordinate regulation of glycolysis by hypoxia in mammalian cells. J Cell Physiol. 1984;118:287–90. doi: 10.1002/jcp.1041180311. [DOI] [PubMed] [Google Scholar]
  • 47.Anderson GR, Stoler DL, Scarcello LA. Normal fibroblasts responding to anoxia exhibit features of the malignant phenotype. J Biol Chem. 1989;264:14885–92. [PubMed] [Google Scholar]
  • 48.Price BD, Calderwood SK. Gadd45 and Gadd153 messenger RNA levels are increased during hypoxia and after exposure of cells to agents which elevate the levels of the glucose-regulated proteins. Cancer Res. 1992;52:3814–7. [PubMed] [Google Scholar]
  • 49.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;92:5510–4. doi: 10.1073/pnas.92.12.5510. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.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;266:15563–6. [PubMed] [Google Scholar]
  • 51.Pugh CW, Tan CC, Jones RW, Ratcliffe PJ. Functional analysis of an oxygen-regulated transcriptional enhancer lying 3′ to the mouse erythropoietin gene. Proc. Natl. Acad. Sci. USA. 1991;88:10553–7. doi: 10.1073/pnas.88.23.10553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Semenza GL, Nejfelt MK, Chi SM, Antonarakis SE. Hypoxia-inducible nuclear factors bind to an enhancer element located 3′ to the human erythropoietin gene. Proc. Natl. Acad. Sci. USA. 1991;88:5680–5684. doi: 10.1073/pnas.88.13.5680. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Madan A, Curtin PT. A 24-base-pair sequence 3′ to the human erythropoietin gene contains a hypoxia-responsive transcriptional enhancer. Proc Natl Acad Sci U S A. 1993;90:3928–32. doi: 10.1073/pnas.90.9.3928. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.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. doi: 10.1128/mcb.12.12.5447. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Wang GL, Semenza GL. Characterization of hypoxia-inducible factor 1 and regulation of DNA binding activity by hypoxia. J. Biol. Chem. 1993;268:21513–21518. [PubMed] [Google Scholar]
  • 56.Jiang BH, Semenza GL, Bauer C, Marti HH. Hypoxia-inducible factor 1 levels vary exponentially over a physiologically relevant range of O2 tension. Am. J. Physiol. 1996;271:C1172–80. doi: 10.1152/ajpcell.1996.271.4.C1172. [DOI] [PubMed] [Google Scholar]
  • 57.Wouters B, Koritzinsky M. Hypoxia signalling through mTOR and the unfolded protein response in cancer. Nature Reviews Cancer. 2008 doi: 10.1038/nrc2501. Epub ahead of print. [DOI] [PubMed] [Google Scholar]
  • 58.Simon MC, Keith B. The role of oxygen availability in embryonic development and stem cell function. Nat Rev Mol Cell Biol. 2008;9:285–96. doi: 10.1038/nrm2354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Jain S, Maltepe E, Lu MM, Simon C, Bradfield CA. Expression of ARNT, ARNT2, HIF1alpha, HIF2alpha and Ah receptor mRNAs in the developing mouse. Mech Dev. 1998;73:117–123. doi: 10.1016/s0925-4773(98)00038-0. [DOI] [PubMed] [Google Scholar]
  • 60.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. doi: 10.1101/gad.11.1.72. [DOI] [PubMed] [Google Scholar]
  • 61.Wiesener MS, et al. Widespread hypoxia-inducible expression of HIF-2alpha in distinct cell populations of different organs. Faseb J. 2003;17:271–273. doi: 10.1096/fj.02-0445fje. [DOI] [PubMed] [Google Scholar]
  • 62.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. [PMC free article] [PubMed] [Google Scholar]
  • 63.Makino Y, et al. Inhibitory PAS domain protein is a negative regulator of hypoxia-inducible gene expression. Nature. 2001;414:550–554. doi: 10.1038/35107085. [DOI] [PubMed] [Google Scholar]
  • 64.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;95:7987–92. doi: 10.1073/pnas.95.14.7987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Salceda S, Caro J. Hypoxia-inducible factor 1alpha (HIF-1alpha) protein is rapidly degraded by the ubiquitin-proteasome system under normoxic conditions. Its stabilization by hypoxia depends on redox-induced changes. J Biol Chem. 1997;272:22642–7. doi: 10.1074/jbc.272.36.22642. [DOI] [PubMed] [Google Scholar]
  • 66.Kallio PJ, Wilson WJ, O’Brien S, Makino Y, Poellinger L. Regulation of the hypoxia-inducible transcription factor 1alpha by the ubiquitin-proteasome pathway. J. Biol. Chem. 1999;274:6519–25. doi: 10.1074/jbc.274.10.6519. [DOI] [PubMed] [Google Scholar]
  • 67.Maxwell PH, et al. The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature. 1999;399:271–275. doi: 10.1038/20459. [DOI] [PubMed] [Google Scholar]
  • 68.Cockman ME, et al. Hypoxia inducible factor-alpha binding and ubiquitylation by the von Hippel-Lindau tumor suppressor protein. J. Biol. Chem. 2000;275:25733–41. doi: 10.1074/jbc.M002740200. [DOI] [PubMed] [Google Scholar]
  • 69.Kamura T, et al. Activation of HIF1alpha ubiquitination by a reconstituted von Hippel-Lindau (VHL) tumor suppressor complex. Proc Natl Acad Sci U S A. 2000;97:10430–5. doi: 10.1073/pnas.190332597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Ohh M, et al. Ubiquitination of hypoxia-inducible factor requires direct binding to the beta-domain of the von Hippel-Lindau protein. Nat. Cell Biol. 2000;2:423–7. doi: 10.1038/35017054. [DOI] [PubMed] [Google Scholar]
  • 71.Tanimura T, Shepard TH. Glucose metabolism by rat embryos in vitro. Proc Soc Exp Biol Med. 1970;135:51–4. doi: 10.3181/00379727-135-34985. [DOI] [PubMed] [Google Scholar]
  • 72.Stebbins CE, Kaelin WG, Jr., Pavletich NP. Structure of the VHL-ElonginC-ElonginB complex: implications for VHL tumor suppressor function. Science. 1999;284:455–61. doi: 10.1126/science.284.5413.455. [DOI] [PubMed] [Google Scholar]
  • 73.Jaakkola P, et al. Targeting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science. 2001;292:468–472. doi: 10.1126/science.1059796. [DOI] [PubMed] [Google Scholar]
  • 74.Ivan M, et al. HIFalpha targeted for VHL-mediated destruction by proline hydroxylation: implications for O2 sensing. Science. 2001;292:464–468. doi: 10.1126/science.1059817. [DOI] [PubMed] [Google Scholar]
  • 75.Yu F, White SB, Zhao Q, Lee FS. HIF-1alpha binding to VHL is regulated by stimulus-sensitive proline hydroxylation. Proc. Natl. Acad. Sci. USA. 2001;98:9630–9635. doi: 10.1073/pnas.181341498. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Epstein AC, et al. C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell. 2001;107:43–54. doi: 10.1016/s0092-8674(01)00507-4. [DOI] [PubMed] [Google Scholar]
  • 77.Bruick RK, McKnight SL. A conserved family of prolyl-4-hydroxylases that modify HIF. Science. 2001;294:1337–1340. doi: 10.1126/science.1066373. [DOI] [PubMed] [Google Scholar]
  • 78.Ivan M, et al. Biochemical purification and pharmacological inhibition of a mammalian prolyl hydroxylase acting on hypoxia-inducible factor. Proc Natl Acad Sci U S A. 2002;99:13459–13464. doi: 10.1073/pnas.192342099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Mahon PC, Hirota K, Semenza GL. FIH-1: a novel protein that interacts with HIF-1alpha and VHL to mediate repression of HIF-1 transcriptional activity. Genes Dev. 2001;15:2675–2686. doi: 10.1101/gad.924501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Hewitson KS, et al. Hypoxia-inducible Factor (HIF) Asparagine Hydroxylase Is Identical to Factor Inhibiting HIF (FIH) and Is Related to the Cupin Structural Family. J Biol Chem. 2002;277:26351–5. doi: 10.1074/jbc.C200273200. [DOI] [PubMed] [Google Scholar]
  • 81.Lando D, et al. FIH-1 is an asparaginyl hydroxylase enzyme that regulates the transcriptional activity of hypoxia-inducible factor. Genes Dev. 2002;16:1466–1471. doi: 10.1101/gad.991402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.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. doi: 10.1126/science.1068592. [DOI] [PubMed] [Google Scholar]
  • 83.Maxwell PH, Pugh CW, Ratcliffe PJ. Activation of the HIF pathway in cancer. Curr Opin Genet Dev. 2001;11:293–9. doi: 10.1016/s0959-437x(00)00193-3. [DOI] [PubMed] [Google Scholar]
  • 84.Semenza GL. Targeting HIF-1 for cancer therapy. Nat Rev Cancer. 2003;3:721–732. doi: 10.1038/nrc1187. [DOI] [PubMed] [Google Scholar]
  • 85.Giaccia A, Siim BG, Johnson RS. HIF-1 as a target for drug development. Nat Rev Drug Discov. 2003;2:803–11. doi: 10.1038/nrd1199. [DOI] [PubMed] [Google Scholar]
  • 86.Birner P, et al. Overexpression of hypoxia-inducible factor 1alpha is a marker for an unfavorable prognosis in early-stage invasive cervical cancer. Cancer Res. 2000;60:4693–6. [PubMed] [Google Scholar]
  • 87.Schindl M, et al. Overexpression of hypoxia-inducible factor 1alpha is associated with an unfavorable prognosis in lymph node-positive breast cancer. Clin Cancer Res. 2002;8:1831–7. [PubMed] [Google Scholar]
  • 88.Bos R, et al. Levels of hypoxia-inducible factor-1 alpha during breast carcinogenesis. J Natl Cancer Inst. 2001;93:309–14. doi: 10.1093/jnci/93.4.309. [DOI] [PubMed] [Google Scholar]
  • 89.Birner P, et al. Expression of hypoxia-inducible factor-1 alpha in oligodendrogliomas: its impact on prognosis and on neoangiogenesis. Cancer. 2001;92:165–71. doi: 10.1002/1097-0142(20010701)92:1<165::aid-cncr1305>3.0.co;2-f. [DOI] [PubMed] [Google Scholar]
  • 90.Birner P, Schindl M, Obermair A, Breitenecker G, Oberhuber G. Expression of hypoxia-inducible factor 1alpha in epithelial ovarian tumors: its impact on prognosis and on response to chemotherapy. Clin Cancer Res. 2001;7:1661–8. [PubMed] [Google Scholar]
  • 91.Sivridis E, Giatromanolaki A, Gatter KC, Harris AL, Koukourakis MI. Association of hypoxia-inducible factors 1alpha and 2alpha with activated angiogenic pathways and prognosis in patients with endometrial carcinoma. Cancer. 2002;95:1055–63. doi: 10.1002/cncr.10774. [DOI] [PubMed] [Google Scholar]
  • 92.Aebersold DM, et al. Expression of hypoxia-inducible factor-1alpha: a novel predictive and prognostic parameter in the radiotherapy of oropharyngeal cancer. Cancer Res. 2001;61:2911–6. [PubMed] [Google Scholar]
  • 93.Giatromanolaki A, et al. Relation of hypoxia inducible factor 1 alpha and 2 alpha in operable non-small cell lung cancer to angiogenic/molecular profile of tumours and survival. Br J Cancer. 2001;85:881–90. doi: 10.1054/bjoc.2001.2018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Holmquist-Mengelbier L, et al. Recruitment of HIF-1alpha and HIF-2alpha to common target genes is differentially regulated in neuroblastoma: HIF-2alpha promotes an aggressive phenotype. Cancer Cell. 2006;10:413–23. doi: 10.1016/j.ccr.2006.08.026. [DOI] [PubMed] [Google Scholar]
  • 95.Khatua S, et al. Overexpression of the EGFR/FKBP12/HIF-2alpha pathway identified in childhood astrocytomas by angiogenesis gene profiling. Cancer Res. 2003;63:1865–70. [PubMed] [Google Scholar]
  • 96.Koukourakis MI, et al. Hypoxia-inducible factor (HIF1A and HIF2A), angiogenesis, and chemoradiotherapy outcome of squamous cell head-and-neck cancer. Int J Radiat Oncol Biol Phys. 2002;53:1192–202. doi: 10.1016/s0360-3016(02)02848-1. [DOI] [PubMed] [Google Scholar]
  • 97.Beasley NJ, et al. Hypoxia-inducible factors HIF-1alpha and HIF-2alpha in head and neck cancer: relationship to tumor biology and treatment outcome in surgically resected patients. Cancer Res. 2002;62:2493–7. [PubMed] [Google Scholar]
  • 98.Volm M, Koomagi R. Hypoxia-inducible factor (HIF-1) and its relationship to apoptosis and proliferation in lung cancer. Anticancer Res. 2000;20:1527–33. [PubMed] [Google Scholar]
  • 99.Mandriota SJ, et al. HIF activation identifies early lesions in VHL kidneys: evidence for site-specific tumor suppressor function in the nephron. Cancer Cell. 2002;1:459–68. doi: 10.1016/s1535-6108(02)00071-5. [DOI] [PubMed] [Google Scholar]
  • 100.Raval RR, 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;25:5675–5686. doi: 10.1128/MCB.25.13.5675-5686.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Maxwell PH, et al. Hypoxia-inducible factor-1 modulates gene expression in solid tumors and influences both angiogenesis and tumor growth. Proc Natl Acad Sci U S A. 1997;94:8104–9. doi: 10.1073/pnas.94.15.8104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Ryan HE, et al. Hypoxia-inducible factor-1alpha is a positive factor in solid tumor growth. Cancer Res. 2000;60:4010–5. [PubMed] [Google Scholar]
  • 103.Kung AL, Wang S, Klco JM, Kaelin WG, Livingston DM. Suppression of tumor growth through disruption of hypoxia-inducible transcription. Nat Med. 2000;6:1335–40. doi: 10.1038/82146. [DOI] [PubMed] [Google Scholar]
  • 104.Liao D, Corle C, Seagroves TN, Johnson RS. Hypoxia-inducible factor-1alpha is a key regulator of metastasis in a transgenic model of cancer initiation and progression. Cancer Res. 2007;67:563–72. doi: 10.1158/0008-5472.CAN-06-2701. [DOI] [PubMed] [Google Scholar]
  • 105.Hiraga T, Kizaka-Kondoh S, Hirota K, Hiraoka M, Yoneda T. Hypoxia and hypoxia-inducible factor-1 expression enhance osteolytic bone metastases of breast cancer. Cancer Res. 2007;67:4157–63. doi: 10.1158/0008-5472.CAN-06-2355. [DOI] [PubMed] [Google Scholar]
  • 106.Koukourakis MI, et al. Endogenous markers of two separate hypoxia response pathways (hypoxia inducible factor 2 alpha and carbonic anhydrase 9) are associated with radiotherapy failure in head and neck cancer patients recruited in the CHART randomized trial. J Clin Oncol. 2006;24:727–35. doi: 10.1200/JCO.2005.02.7474. [DOI] [PubMed] [Google Scholar]
  • 107.Graeber TG, et al. Hypoxia-mediated selection of cells with diminished apoptotic potential in solid tumours. Nature. 1996;379:88–91. doi: 10.1038/379088a0. see comments. [DOI] [PubMed] [Google Scholar]
  • 108.An WG, et al. Stabilization of wild-type p53 by hypoxia-inducible factor 1alpha. Nature. 1998;392:405–8. doi: 10.1038/32925. [DOI] [PubMed] [Google Scholar]
  • 109.Koshiji M, et al. HIF-1alpha induces cell cycle arrest by functionally counteracting Myc. Embo J. 2004;23:1949–1956. doi: 10.1038/sj.emboj.7600196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Gordan JD, Bertout JA, Hu CJ, Diehl JA, Simon MC. HIF-2alpha promotes hypoxic cell proliferation by enhancing c-myc transcriptional activity. Cancer Cell. 2007;11:335–347. doi: 10.1016/j.ccr.2007.02.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Gordan JD, Thompson CB, Simon MC. HIF and c-Myc: sibling rivals for control of cancer cell metabolism and proliferation. Cancer Cell. 2007;12:108–13. doi: 10.1016/j.ccr.2007.07.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Schwarz G. Ueber Desensibilisierung gegen rontgen- und radiumstrahlen. Munchener Medizinische Wochenschrift. 1909;24:1–2. [Google Scholar]
  • 113.Brown JM. Tumor hypoxia in cancer therapy. Methods in Enzymology. 2007:435. doi: 10.1016/S0076-6879(07)35015-5. [DOI] [PubMed] [Google Scholar]
  • 114.Overgaard J. Hypoxic radiosensitization: adored and ignored. J Clin Oncol. 2007;25:4066–74. doi: 10.1200/JCO.2007.12.7878. [DOI] [PubMed] [Google Scholar]
  • 115.Urtasun RC, Band PR, Chapman JD, Feldstein ML. Radiation plus metronidazole for glioblastoma. N Engl J Med. 1977;296:757. doi: 10.1056/NEJM197703312961315. [DOI] [PubMed] [Google Scholar]
  • 116.Zeman EM, Brown JM, Lemmon MJ, Hirst VK, Lee WW. SR-4233: a new bioreductive agent with high selective toxicity for hypoxic mammalian cells. Int J Radiat Oncol Biol Phys. 1986;12:1239–42. doi: 10.1016/0360-3016(86)90267-1. [DOI] [PubMed] [Google Scholar]
  • 117.Brown JM, Wilson WR. Exploiting tumour hypoxia in cancer treatment. Nat Rev Cancer. 2004;4:437–47. doi: 10.1038/nrc1367. [DOI] [PubMed] [Google Scholar]
  • 118.Rischin D, et al. Tirapazamine, Cisplatin, and Radiation versus Fluorouracil, Cisplatin, and Radiation in patients with locally advanced head and neck cancer: a randomized phase II trial of the Trans-Tasman Radiation Oncology Group (TROG 98.02) J Clin Oncol. 2005;23:79–87. doi: 10.1200/JCO.2005.01.072. [DOI] [PubMed] [Google Scholar]
  • 119.von Pawel J, et al. Tirapazamine plus cisplatin versus cisplatin in advanced non-small-cell lung cancer: A report of the international CATAPULT I study group. Cisplatin and Tirapazamine in Subjects with Advanced Previously Untreated Non-Small-Cell Lung Tumors. J Clin Oncol. 2000;18:1351–9. doi: 10.1200/JCO.2000.18.6.1351. [DOI] [PubMed] [Google Scholar]
  • 120.Shrieve DC, Harris JW. The in vitro sensitivity of chronically hypoxic EMT6/SF cells to X-radiation and hypoxic cell radiosensitizers. Int J Radiat Biol Relat Stud Phys Chem Med. 1985;48:127–38. doi: 10.1080/09553008514551131. [DOI] [PubMed] [Google Scholar]
  • 121.Graeber TG, et al. Hypoxia-mediated selection of cells with diminished apoptotic potential in solid tumours. Nature. 1996;379:88–91. doi: 10.1038/379088a0. [DOI] [PubMed] [Google Scholar]
  • 122.Wartenberg M, et al. Regulation of the multidrug resistance transporter P-glycoprotein in multicellular tumor spheroids by hypoxia-inducible factor (HIF-1) and reactive oxygen species. Faseb J. 2003;17:503–5. doi: 10.1096/fj.02-0358fje. [DOI] [PubMed] [Google Scholar]
  • 123.Comerford KM, et al. Hypoxia-inducible factor-1-dependent regulation of the multidrug resistance (MDR1) gene. Cancer Res. 2002;62:3387–94. [PubMed] [Google Scholar]
  • 124.Koshiji M, et al. HIF-1alpha induces genetic instability by transcriptionally downregulating MutSalpha expression. Mol Cell. 2005;17:793–803. doi: 10.1016/j.molcel.2005.02.015. [DOI] [PubMed] [Google Scholar]
  • 125.Erler JT, et al. Lysyl oxidase is essential for hypoxia-induced metastasis. Nature. 2006;440:1222–6. doi: 10.1038/nature04695. [DOI] [PubMed] [Google Scholar]
  • 126.Staller P, et al. Chemokine receptor CXCR4 downregulated by von Hippel-Lindau tumour suppressor pVHL. Nature. 2003;425:307–11. doi: 10.1038/nature01874. [DOI] [PubMed] [Google Scholar]
  • 127.Pennacchietti S, et al. Hypoxia promotes invasive growth by transcriptional activation of the met protooncogene. Cancer Cell. 2003;3:347–61. doi: 10.1016/s1535-6108(03)00085-0. [DOI] [PubMed] [Google Scholar]
  • 128.Moeller BJ, Richardson RA, Dewhirst MW. Hypoxia and radiotherapy: opportunities for improved outcomes in cancer treatment. Cancer Metastasis Rev. 2007;26:241–8. doi: 10.1007/s10555-007-9056-0. [DOI] [PubMed] [Google Scholar]
  • 129.Fujibayashi Y, et al. Copper-62-ATSM: a new hypoxia imaging agent with high membrane permeability and low redox potential. J Nucl Med. 1997;38:1155–60. [PubMed] [Google Scholar]
  • 130.Evans SM, et al. Noninvasive detection of tumor hypoxia using the 2-nitroimidazole [18F]EF1. J Nucl Med. 2000;41:327–36. [PubMed] [Google Scholar]
  • 131.Rasey JS, et al. Characterization of radiolabeled fluoromisonidazole as a probe for hypoxic cells. Radiat Res. 1987;111:292–304. [PubMed] [Google Scholar]
  • 132.Majumder PK, et al. mTOR inhibition reverses Akt-dependent prostate intraepithelial neoplasia through regulation of apoptotic and HTF-1-dependent pathways. Nat Med. 2004;10:594–601. doi: 10.1038/nm1052. [DOI] [PubMed] [Google Scholar]
  • 133.Thomas GV, et al. Hypoxia-inducible factor determines sensitivity to inhibitors of mTOR in kidney cancer. Nat Med. 2006;12:122–7. doi: 10.1038/nm1337. [DOI] [PubMed] [Google Scholar]
  • 134.Liu YV, et al. RACK1 Competes with HSP90 for Binding to HTF-1alpha and Is Required for O(2)-Independent and HSP90 Inhibitor-Induced Degradation of HIF-1alpha. Mol Cell. 2007;25:207–17. doi: 10.1016/j.molcel.2007.01.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Kung AL, et al. Small molecule blockade of transcriptional coactivation of the hypoxia-inducible factor pathway. Cancer Cell. 2004;6:33–43. doi: 10.1016/j.ccr.2004.06.009. [DOI] [PubMed] [Google Scholar]
  • 136.Semenza GL. Evaluation of HIF-1 inhibitors as anticancer agents. Drug Discov Today. 2007;12:853–9. doi: 10.1016/j.drudis.2007.08.006. [DOI] [PubMed] [Google Scholar]
  • 137.Ferrara N, Hillan KJ, Gerber HP, Novotny W. Discovery and development of bevacizumab, an anti-VEGF antibody for treating cancer. Nat Rev Drug Discov. 2004;3:391–400. doi: 10.1038/nrd1381. [DOI] [PubMed] [Google Scholar]
  • 138.Hurwitz H, et al. Bevacizumab plus irinotecan, fluorouracil, and leucovorin for metastatic colorectal cancer. N Engl J Med. 2004;350:2335–42. doi: 10.1056/NEJMoa032691. [DOI] [PubMed] [Google Scholar]
  • 139.Escudier B, et al. Sorafenib in advanced clear-cell renal-cell carcinoma. N Engl J Med. 2007;356:125–34. doi: 10.1056/NEJMoa060655. [DOI] [PubMed] [Google Scholar]
  • 140.Ellis LM, Hicklin DJ. VEGF-targeted therapy: mechanisms of anti-tumour activity. Nat Rev Cancer. 2008 doi: 10.1038/nrc2403. [DOI] [PubMed] [Google Scholar]
  • 141.Paez JG, et al. EGFR mutations in lung cancer: correlation with clinical response to gefitinib therapy. Science. 2004;304:1497–500. doi: 10.1126/science.1099314. [DOI] [PubMed] [Google Scholar]
  • 142.Lynch TJ, et al. Activating mutations in the epidermal growth factor receptor underlying responsiveness of non-small-cell lung cancer to gefitinib. N Engl J Med. 2004;350:2129–39. doi: 10.1056/NEJMoa040938. [DOI] [PubMed] [Google Scholar]
  • 143.Kulshreshtha R, Davuluri RV, Calin GA, Ivan M. A microRNA component of the hypoxic response. Cell Death Differ. 2008;15:667–71. doi: 10.1038/sj.cdd.4402310. [DOI] [PubMed] [Google Scholar]
  • 144.Percy MJ, et al. Novel exon 12 mutations in the HIF2A gene associated with erythrocytosis. Blood. 2008;111:5400–2. doi: 10.1182/blood-2008-02-137703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Percy MJ, et al. A gain-of-function mutation in the HIF2A gene in familial erythrocytosis. N Engl J Med. 2008;358:162–8. doi: 10.1056/NEJMoa073123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.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;117:3879–89. doi: 10.1172/JCI32614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Jurgensen JS, et al. Persistent induction of HIF-1alpha and -2alpha in cardiomyocytes and stromal cells of ischemic myocardium. Faseb J. 2004;18:1415–7. doi: 10.1096/fj.04-1605fje. [DOI] [PubMed] [Google Scholar]
  • 148.Loor G, Schumacker PT. Role of hypoxia-inducible factor in cell survival during myocardial ischemia-reperfusion. Cell Death Differ. 2008;15:686–90. doi: 10.1038/cdd.2008.13. [DOI] [PubMed] [Google Scholar]
  • 149.Kido M, et al. Hypoxia-inducible factor 1-alpha reduces infarction and attenuates progression of cardiac dysfunction after myocardial infarction in the mouse. J Am Coll Cardiol. 2005;46:2116–24. doi: 10.1016/j.jacc.2005.08.045. [DOI] [PubMed] [Google Scholar]
  • 150.Yu AY, et al. Impaired physiological responses to chronic hypoxia in mice partially deficient for hypoxia-inducible factor 1alpha. J Clin Invest. 1999;103:691–6. doi: 10.1172/JCI5912. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Brusselmans K, et al. Heterozygous deficiency of hypoxia-inducible factor-2alpha protects mice against pulmonary hypertension and right ventricular dysfunction during prolonged hypoxia. J Clin Invest. 2003;111:1519–27. doi: 10.1172/JCI15496. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Cramer T, et al. HIF-1alpha is essential for myeloid cell-mediated inflammation. Cell. 2003;112:645–57. doi: 10.1016/s0092-8674(03)00154-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Karhausen J, et al. Epithelial hypoxia-inducible factor-1 is protective in murine experimental colitis. J Clin Invest. 2004;114:1098–106. doi: 10.1172/JCI21086. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES