Open access peer-reviewed chapter

HIF Pathways in Clear Cell Renal Cancer

Written By

Olivia Lombardi and David Robert Mole

Submitted: 01 February 2021 Reviewed: 09 February 2021 Published: 19 March 2021

DOI: 10.5772/intechopen.96539

From the Edited Volume

Biomarkers and Bioanalysis Overview

Edited by Ane Claudia Fernandes Nunes

Chapter metrics overview

519 Chapter Downloads

View Full Metrics

Abstract

Clear cell renal cancers (ccRCC) are characterized by inactivation of the VHL (von Hippel–Lindau) tumor suppressor. Work leading to the 2019 Nobel Prize for Physiology or Medicine has shown that this is central to cellular oxygen-sensing, orchestrated by the HIF (hypoxia-inducible factor) transcription factors. These regulate hundreds of genes that underpin many hallmarks of cancer, including angiogenesis, cellular energetics, cell proliferation, resisting cell death, and avoiding immune destruction. However, HIF also promotes processes that are detrimental to cancer cells. Therefore, the overall consequence of HIF pathway activation is a balance of these influences. We explore how variations in the HIF pathway during tumorigenesis alter this balance to promote ccRCC formation.

Keywords

  • cancer
  • kidney
  • renal
  • clear cell
  • von Hippel Lindau
  • VHL
  • hypoxic
  • hypoxia-inducible factor
  • HIF

1. Introduction

Kidney cancer is the seventh most common malignancy in the Western world. In 2018, there were approximately 400,000 new kidney cancer cases and 180,000 kidney cancer-associated deaths worldwide [1]. The underlying causes of kidney cancer are complex and incompletely understood, although genetic factors (both inherited and somatic genetic mutations) are known to drive the disease. Additionally, certain lifestyle choices (such as smoking and a high protein diet) increase the risk of developing kidney cancer, consistent with its prevalence in the Western world. Unless surgically resectable, kidney cancer is largely incurable, and the 5-year survival rate for those with metastatic disease is only about 10% [2]. Systemic anti-cancer therapies, including those that inhibit the vascular response or enhance patients’ immune response to the malignancy, have offered some hope [3]. However, these treatments confer limited efficacy and a considerable burden of toxicity. Therefore, there is a pressing need to better understand the drivers of kidney cancer in order to identify novel therapeutic strategies.

Advertisement

2. Histological subtypes of renal cancers

The most common form of kidney cancer is clear cell renal cell carcinoma (ccRCC), which arises from the adult renal tubular epithelium and accounts for approximately 75% of all kidney cancer cases. This subtype is termed as such due to the characteristic ‘clear’ cytoplasm of malignant cells observed histologically. This is caused by accumulation of excess glycogen and lipid in the cytoplasm (due to highly dysregulated metabolic pathways), which are dissolved by the tissue fixation process [4]. Other less common subtypes of adult renal cancers that also arise from the tubular epithelium include papillary RCC (types 1 and 2); chromophobe RCC; and oncocytoma. Each subtype is associated with different histological features, genetic drivers, and clinical behaviors. Rarely, cancers can arise from other cell types in the adult kidney, including transitional epithelial cells of the ureter and renal pelvis (giving rise to transitional cell carcinoma) and various mesenchymal cell types (e.g. interstitial cells, giving rise to renomedullary interstitial cell tumors). Although childhood kidney cancer is generally rare, the most common form is Wilms tumor, which originates in developing tubular cells during fetal development [5].

It should be noted that even within a specific renal cancer histological subtype there is evidence for substantial heterogeneity, which has initiated efforts to further refine subtype classification based on additional features. Recent studies have found that ccRCC can be further stratified based on architectural, cytological and microenvironmental features, and that these features can predict patient outcome and response to therapy [6]. The underlying cause of this variation remains to be determined but could be due to certain genetic or epigenetic differences between ccRCC tumors. Consolidation of histological and molecular heterogeneity in ccRCC will be important for disease subclassification, as well as better understanding ccRCC biology, going forward.

Advertisement

3. VHL syndrome

Each kidney cancer subtype is associated with its own monogenic cancer syndrome [7]. Studying these rare family kindreds has provided unique insight into the genetic mechanisms underlying both inherited and sporadic cancers. In particular, clear cell renal cancer is associated with VHL syndrome, which is an autosomal dominant disorder, affecting 1 in 32,000 individuals, caused by heterozygous germline mutations of the VHL gene [8, 9]. As well as ccRCC, VHL syndrome is associated with a limited number of other tumors types, including hemangioblastomas of the retina and the central nervous system; pheochromocytomas; pancreatic lesions; endolymphatic sac tumors and epidydimal cysts [8, 9]. VHL syndrome can be further sub-divided according to which of these different tumor types develop in individuals within the kindred [10, 11]. Four distinct patterns have been identified: type 1 VHL disease, which is associated with hemangioblastoma and ccRCC; type 2A, which is associated with hemangioblastoma and pheochromocytoma; type 2B, which is associated with hemangioblastoma, pheochromocytoma and ccRCC; and type 2C, which is associated with pheochromocytoma alone. Each of these subtypes is linked to particular types of VHL mutation, which have been shown to have different downstream biological effects [12, 13, 14, 15, 16, 17].

Advertisement

4. The VHL gene

The human VHL gene was first identified following classical linkage analysis of families with VHL syndrome and was cloned in 1993 [18]. In humans it is located on the short arm of chromosome 3 (3p25) and has three exons that encode a protein of 213 amino acids, with a molecular weight of around 30 kDa (termed p30). However, the gene also contains a second translation start site at codon 53, leading to the generation of a shorter protein of approximately 19 kDa (termed p19), which appears to retain canonical activity [19]. As a consequence, oncogenic mutations, most typically single-nucleotide variants (SNVs) or short insertion/deletions (indels), are restricted to codons 53–213 in exons two and three.

VHL acts as a tumor suppressor gene [20, 21]. ccRCC and other cancer types are associated with inactivating mutations of VHL, which lead to loss-of-function of the gene product (termed pVHL). Although autosomal dominant at the level of the individual, both alleles of the VHL gene must be inactivated in a cell for cancer to develop, in line with Knudson’s two-hit hypothesis [22, 23]. Since VHL syndrome is caused by germline VHL mutation, all cells of the affected individual harbor this mutation. The remaining wild-type (WT) allele is somatically inactivated in the tumor progenitor cell, which then multiplies to form the cancer [20, 21]. Typically, somatic inactivation of the WT allele occurs as a result of an arm-level loss of chromosome 3p (Figure 1), although promoter hypermethylation or a second SNV/indel may also cause complete loss of functional VHL in the cell. Furthermore, since the cells of patients with VHL syndrome only require one somatic mutation to become functionally deficient in VHL, it is a relatively common event, accounting for the high tumor penetrance in these individuals. Indeed, over the course of their lifetime, these individuals often develop multiple tumors and close examination of their organs often reveals the presence of numerous synchronous tumors. However, VHL mutation is only associated with the very limited range of cancers outlined above, despite it being ubiquitously expressed. Therefore, VHL only appears to act as a tumor suppressor gene in very few tissues. Indeed, even within the kidney, ccRCCs appear to develop from a subset of proximal tubular cells [24]. It is assumed that somatic mutations in the wild-type copy of VHL do occur in other cell types, but it is not known whether these cells are eliminated by other tumor suppressor mechanisms, or simply fail to progress to overt cancer.

Figure 1.

VHL inactivation in ccRCC. Individuals with VHL syndrome are predisposed to ccRCC (termed hereditary ccRCC) as a result of a heterozygous germline VHL mutation. The second, wild-type allele is subsequently inactivated by somatic loss of chromosome 3p, resulting in biallelic VHL inactivation. On the other hand, in sporadic ccRCC, two somatic events are required for biallelic inactivation. Typically, one copy of chromosome 3p is lost followed by inactivation of the second VHL allele through mutation or promoter hypermethylation. Although the ordering is reversed, the same genetic aberrations are observed in both sporadic and hereditary ccRCC. However, because only one somatic event is required for biallelic VHL inactivation in patients with VHL syndrome, this is a much more likely event and occurs in multiple cells within the kidney, causing many pre-malignant lesions and multiple ccRCC tumors. chr= chromosome; CNAs= chromosomal copy number alterations; mut= mutation; WT= wild-type.

Importantly, VHL is also inactivated in the vast majority (approximately 90%) of sporadic ccRCC tumors, which occur in patients without a germline mutation in the VHL gene [25]. In order to develop cancer, these individuals require two somatic events affecting both copies of the VHL gene in the same cell (Figure 1). As a result, this occurs much less frequently, accounting for the much lower overall prevalence of ccRCC in the general population of about 1%. However, in contrast to VHL syndrome, the order of events is typically reversed, with loss of chromosome 3p frequently occurring first and the remaining copy then being inactivated by single-nucleotide substitution (SNV) or small insertions or deletions (indels) [26].

Of note, although biallelic VHL inactivation is required for ccRCC (and other tumors) to develop, it does not appear to be sufficient on its own (Figure 1). Mitchell et al. have estimated that in sporadic ccRCC, VHL inactivation predates tumor formation by a number of years or even decades [26]. Consistent with this, examination of the kidneys from patients with VHL syndrome has identified multiple isolated VHL-defective cells, which may be present as single cells or small non-invasive cysts [27, 28]. Furthermore, in vitro, inactivation of VHL leads to cellular senescence rather than unrestricted proliferation [29, 30]. Therefore, it is thought that additional gene mutations are required for these early VHL-defective lesions to develop into mature ccRCC. Indeed, more recently, additional somatic mutations have been identified in ccRCC [25, 31, 32, 33]. Most notable among these are inactivating mutations in the PBRM1 (polybromo 1), SETD2 (SET domain-containing 2) and BAP1 (BRCA-associated protein 1) tumor suppressor genes, mutation of which typically follows loss of VHL. Importantly, these three genes also reside on the short arm of chromosome 3. As a result, the loss of chromosome 3p frequently observed in both familial and sporadic ccRCC can simultaneously result in copy loss of all 4 of these ccRCC-associated tumor suppressor genes; VHL, PBRM1, SETD2 and BAP1.

Advertisement

5. Function of pVHL

Following identification and cloning of the VHL tumor suppressor gene, its sequence did not immediately suggest a function for the protein. However, early immunoprecipitation experiments indicated that pVHL forms a complex with elongin B and elongin C [34]; cullin 2, a member of the Cdc53 family of proteins [35]; and the RING-box protein Rbx1 [36, 37]. Importantly, the binding of pVHL to elongins B and C could be blocked by specific ccRCC-associated mutations in the VHL gene, strongly suggesting that these two proteins contribute to the tumor suppressor activity of VHL [34]. The subsequent identification of mutations in the TCEB1 gene, which encodes elongin C, in ccRCC tumors that have wild-type VHL further emphasizes the importance of this complex in ccRCC formation [25, 32, 38].

Elongins B and C, cullin 2 and Rbx1 are all components of an E3-ligase complex that adds polyubiquitin chains to specific proteins and thus targets them for degradation by the proteasome [39, 40]. This suggested that pVHL might act as the recognition component of a pVHL ligase complex. In a separate line of work, dysregulation of the hypoxia-inducible factor (HIF) transcription factors had been identified in VHL-defective ccRCC cells [41]. It was subsequently shown that pVHL directly interacted with HIF, leading to polyubiquitination and subsequent proteasomal degradation of its alpha-subunits [42, 43]. Again, the pVHL-HIF interaction could be blocked by specific ccRCC-associated mutations in VHL, leading to overexpression of HIF and underlining the importance of HIF in the development of ccRCC [42]. Importantly, this interaction was not only altered by pathogenic VHL mutations but was also regulated in an oxygen-dependent manner [44, 45]. This indicated that the pVHL-HIF interaction was integral to the mechanism of cellular oxygen-sensing.

The central role of HIF in ccRCC biology has been further underscored in numerous studies. In particular, in xenograft and transgenic mouse models of VHL-defective ccRCC, tumor growth is dependent upon the presence of HIF [46, 47, 48, 49, 50, 51]. Specifically, tumor growth is dependent on the DNA binding activity of HIF, which is required for it to transactivate its target genes [48]. Thus, HIF and its associated transcriptional response are key mediators of tumorigenesis in ccRCC.

In addition to HIF, pVHL can interact with a number of other proteins, although the biological significance of these interactions is incompletely understood [52]. Some of these interactions can lead to ubiquitination of other proteins aside from HIF. For example, pVHL has been reported to interact with and ubiquitinate two de-ubiquitinase enzymes (VDU1 and VDU2) leading to their degradation [53, 54]. In turn, VDU2 but not VDU1 may de-ubiquitinate HIF-1α, potentially providing another level of control to the HIF pathway [55]. In addition, pVHL can bind to and ubiquitinate two subunits of the RNA polymerase 2 complex, POL2RA (RPB1) and POL2RG (RPB7) [56, 57, 58]. Importantly, the pVHL-RPB1 interaction was shown to be oxygen-dependent, involving a mechanism similar to that regulating pVHL interaction with HIF [58]. Similarly, the erythropoietin receptor (EPOR), which lies downstream of the canonical HIF-target gene, erythropoietin (EPO), may also be bound and ubiquinated by pVHL in response to oxygen [59]. pVHL can also interact with and ubiquitinate the regulatory domain of atypical protein kinase C (PKC), a serine–threonine kinase that has roles in cell polarity and cell growth, leading to its degradation [60, 61, 62]. Again, this interaction may be regulated by oxygen [62]. Similarly, an oxygen-dependent interaction between pVHL and sprouty homolog 2 (SPRY2), which modulates the action of receptor tyrosine kinases, has been reported [63]. Taken together, these findings indicate that pVHL may contribute to oxygen signaling more extensively than simply through regulation of HIF.

pVHL may also play a non-canonical role in extra-cellular matrix assembly, independently of HIF. Specifically, pVHL can interact directly with the alpha-chain of collagen 4 and is important in maintaining the collagen 4 network [64, 65]. This molecule is heavily hydroxylated, and as will be explained below, hydroxylation is important in the recognition of HIF-alpha (as well as collagen 4) by pVHL. Importantly, this interaction can be dissociated by ccRCC-associated VHL mutations. Similarly, fibronectin co-immunoprecipitates with pVHL, and consistently the extracellular fibronectin matrix produced by VHL-defective ccRCC cells is also disrupted [66]. However, the contribution of this phenomenon to cellular oxygen sensing and ccRCC tumorigenesis is still unclear.

Advertisement

6. Oxygen-dependent regulation of HIF by pVHL

The importance of pVHL in the regulation of the HIF transcription factors, and the cellular transcriptional response to altered levels of oxygen, has provided tremendous insights into the mechanisms of cellular oxygen sensing. HIF was first discovered in the quest for transcriptional regulators of the erythropoietin gene (EPO), encoding the master regulator of red blood cell production [67]. It later emerged there were three HIF isoforms, HIF-1, HIF-2 [68, 69], and HIF-3 [70], each composed of a common, constitutive β-subunit (HIF-1β, also known as ARNT – aryl hydrocarbon receptor nuclear translocator) and a regulated alpha-subunit (HIF-1α, HIF-2α and HIF-3α respectively). HIF-1α is ubiquitously expressed at the mRNA level, thus HIF-1α protein is capable of being stabilized in all tissue types. HIF-1α is thought to drive core, canonical cellular responses to low oxygen levels (hypoxia) [71], including the metabolic switch to anaerobic glycolysis. The expression of HIF-2α mRNA and HIF-3α mRNA is more cell-type-specific and thus these transcription factors are thought to drive more specialized responses to hypoxia [69, 70, 72]. HIF-2α expression is generally more restricted to particular mesenchymal cell types, including endothelial cells in which it was first identified, hence its alias endothelial PAS domain-containing protein 1 (EPAS1) [69]. However, HIF-2α is also expressed in some epithelial malignancies, including ccRCC. HIF-3α expression is restricted to a select few cell types and can be alternatively spliced to yield several transcript variants [70]. The biological functions of HIF-3α have not been well-explored, although it is thought to antagonize the transcriptional responses of HIF-1α and HIF-2α [73, 74, 75].

HIF isoforms are all basic helix–loop–helix/Per-ARNT-SIM (bHLH–PAS) transcription factors, belonging to a much larger family that includes the oncogenic MYC proteins [76]. Each possess an N-terminal bHLH DNA-binding domain and two protein–protein interaction PAS domains responsible for dimerization. In addition, the three HIF-α isoforms each contain oxygen-dependent degradation domains (ODDDs), responsible for regulating protein abundance [77]. However, only HIF-1α and HIF-2α possess the C-terminal transactivation domains (C-TAD) [78].

In the presence of oxygen, HIF-α subunits are hydroxylated on two residues in the ODDD domains by a family of prolyl hydroxylase enzymes (PHD1, PHD2 and PHD3) [44, 45, 79]. These hydroxylated residues are recognized and bound by pVHL (in a complex with elongin B, elongin C and cullin 2) leading to its rapid ubiquitination and proteasomal degradation (Figure 2). Thus, when oxygen is abundant, HIF-α levels are low. However, since oxygen is a rate-limiting substrate for this reaction, HIF-α is stabilized in hypoxia. Inactivation of VHL in ccRCC cells will also block HIF from being degraded, leading to constitutive activation of HIF and its target genes, even in cells that are well-oxygenated. Accordingly, activation of both HIF and HIF target genes are hallmarks of ccRCC.

Figure 2.

Regulation of HIF by PHD enzymes and pVHL E3 ligase. (A) In normal oxygen conditions (normoxia), the oxygen-dependent PHD enzymes (PHD1, PHD2 and PHD3) hydroxylate both HIF-1α and HIF-2α transcription factor isoforms. This causes HIF proteins to be recognized and ubiquitinated by the pVHL E3 ubiquitin ligase complex, which targets them for rapid degradation via the proteasome. (B) In low oxygen conditions (hypoxia), PHD enzymes are inactive due to the lack of their oxygen substrate. Therefore, HIF-1α and HIF-2α are not hydroxylated and are not targeted for degradation by pVHL. Due to their stabilization, they are able to dimerize with their obligate binding partner HIF-1β. This allows them to bind to DNA and upregulate their target genes. (C) When the VHL gene is inactivated (as is the case in ccRCC and some other cancers), pVHL is either not expressed or is dysfunctional. Therefore, pVHL is unable to recognize HIF-1α and HIF-2α, even in the presence of oxygen when they are hydroxylated by PHD enzymes. This causes inappropriate stabilization of HIF-1α and HIF-2α, which then dimerize with HIF-1β and upregulate their target genes, regardless of oxygen levels.

In addition, HIF-1α and HIF-2α can be further modified at an additional site in the C-terminal TAD by an asparaginyl hydroxylase, termed factor inhibiting HIF 1 (FIH-1) [80, 81]. Similar to the PHD enzymes, FIH-1 activity is oxygen-dependent, but asparagine hydroxylation does not prompt recognition by pVHL. Instead, asparaginyl hydroxylated HIF is unable to bind to the transcriptional co-factor, CREB binding protein (CBP)/p300, which facilitates transcriptional activation at a subset of HIF-target genes [80, 81, 82]. Therefore, two distinct mechanisms act to control HIF activity and expression in an oxygen-dependent manner, one of which is blocked by VHL inactivation. In the context of ccRCC, this has two consequences. Firstly, FIH-1 may facilitate residual hypoxic regulation of HIF despite constitutive HIF stabilization [83]. Secondly, the transcriptional response to VHL inactivation in normoxic cells may not precisely mimic the transcriptional response to hypoxia.

Advertisement

7. The HIF transcriptional response

Once stabilized, both HIF-1α and HIF-2α, in complex with HIF-1β, are able to bind chromatin at either gene promoters or promoter-distant enhancers that contain one or more 5′-RCGTG-3′ recognition motifs, termed hypoxia response elements (HREs) [84, 85]. These short motifs are highly numerous across the genome and only a small proportion of accessible motifs are occupied by HIF, indicating that additional factors are involved in HIF DNA-binding [85]. HIF-binding sites may lie several hundreds of kilobases from the target promoter, interacting with it through chromatin looping, which can make it difficult to identify the transcriptional target of any given binding site. Therefore, much effort has been directed at determining both direct and indirect targets of the HIF transcriptional pathway in multiple settings, including in VHL-defective ccRCC cells, using both transcriptomic assays such as RNA-seq and assays of chromatin binding such as ChIP-seq [85, 86, 87, 88, 89].

These sequencing studies indicate that HIF acts as a gene activator rather than a repressor; causing the induction of hundreds to thousands of genes and triggering massive pathway activation [90, 91, 92, 93]. These genes mediate diverse cellular functions including angiogenesis, erythropoiesis, glycolysis and the cell cycle [77, 94, 95]. This triggers a physiological response that enables cells to survive in low oxygen conditions. For example, HIF-dependent angiogenesis increases blood supply to oxygen-starved tissue; HIF-dependent erythropoiesis improves systemic oxygen delivery; HIF-dependent glycolysis allows cells to generate ATP in the absence of oxygen; and HIF-dependent cell cycle arrest can allow cells to conserve energy and reduce oxygen consumption.

Importantly, HIF-binding sites and HIF-regulated genes are highly cell-type specific. Thus, whilst HIF may regulate many hundreds of genes in any given cell type, only a small, core set of well-described genes are regulated in the majority of tissues [90, 93]. Furthermore, although both HIF-1α and HIF-2α share the same binding motif and their binding sites often overlap, HIF-1α tends to be more prevalent at gene promoters whereas HIF-2α is more prevalent at promoter-distant enhancers [90, 92]. In addition to this binding site specificity, post-DNA-binding mechanisms likely contribute to transcriptional selectivity between the two isoforms [96], such that specific genes may be regulated by either HIF-1α or HIF-2α only, even when both isoforms are bound [50, 97] For example, cyclin D1 (CCND1), transforming growth factor alpha (TGFA), vascular endothelial growth factor A (VEGFA), glucose uptake transporter 1 (SLC2A1/GLUT1), the MYC oncogene, and the stemness-related transcription factor OCT4/POU5F1 are specifically induced by HIF-2, whilst BCL2-interacting protein 3 (BNIP3) and carbonic anhydrase 9 (CA9) are positively regulated by HIF-1 [97, 98, 99, 100, 101, 102].

Although primarily a physiological response, the HIF pathway is also relevant to the pathophysiology of cancer and many HIF target genes are central to the hallmarks of cancer described by Hanahan and Weinberg [103]. These include genes with prominent roles in angiogenesis, glycolysis, cell proliferation, cell invasion and immune evasion among other oncogenic processes (Figure 3). Indeed, HIF is activated in many types of solid tumor, largely as a result of intra-tumor hypoxia and is almost universally associated with a poor prognosis [104].

Figure 3.

HIF target genes that promote or restrict tumorigenesis. HIF regulates hundreds to thousands of target genes, which mediate diverse and sometimes conflicting cellular processes. For example, such processes can either promote or restrict tumor growth. Those that are typically considered tumor-promoting processes are depicted in red, whereas those that are typically considered tumor-suppressive are depicted in green. Cellular processes that can be either tumor-promoting or -suppressive (depending on the context) are depicted in red and green. Exemplar HIF target genes involved in each process are listed. Note that whilst some HIF target genes appear to be consistent across cell types and conditions, others are context-dependent.

In particular, HIF promotes the metabolic switch from oxidative phosphorylation to anaerobic glycolysis by inducing a range of target genes, including those encoding transmembrane proteins that import glucose into the cell (SLC2A1/GLUT-1 and SLC2A3/GLUT-3) as well as multiple catalytic enzymes in the glycolytic pathway [71]. Oxidative phosphorylation is oxygen-dependent, therefore switching to oxygen-independent glycolysis allows hypoxic cancer cells to generate energy. However, glycolysis causes accumulation of byproducts in the form of acidic metabolites, which can be toxic to cancer cells. Therefore, HIF also upregulates genes encoding transmembrane proteins that rebalance intracellular pH to promote cancer cell survival. For example, the HIF target genes CA9 and CA12, encoding carbonic anhydrases, generate alkaline sodium bicarbonate ions in the extracellular space [105]. Sodium bicarbonate can then be imported into cells by ion channels to counteract intracellular acidity. Furthermore, once a tumor outgrows its blood supply and becomes hypoxic, HIF induces genes encoding pro-angiogenic secreted factors, such as VEGFA and placental growth factor (PGF), that serve to transmit extracellular signals and stimulate blood vessel production [106]. This increases delivery of nutrients and oxygen to cancer cells, enabling the tumor to further expand. Furthermore, HIF has recently been found to upregulate genes that help cancer cells evade destruction by the immune system. One such example is CD274, encoding the transmembrane protein termed programed death ligand 1 (PD-L1), which is expressed in cancer cells [107]. PD-L1 interacts with its receptor termed programed cell death protein 1 (PD-1), which is expressed on the cell surface of T cells. The PD-L1/PD-1 interaction prevents T cell-mediated killing of cancer cells, therefore HIF may exacerbate this oncogenic mechanism.

However, since HIF evolved to mediate physiological responses to hypoxia, not all HIF target genes are advantageous in a cancer setting. Paradoxically, although HIF activates many pro-tumorigenic target genes, there are also anti-tumorigenic HIF targets (Figure 3). These may represent in-built tumor suppressor mechanisms that counterbalance oncogenic target genes when HIF is activated in response to physiological hypoxia. Tumor suppressive HIF target genes include BNIP3 and BNIP3L, which are pro-apoptotic proteins. BNIP3 and BNIP3L can promote either cell death or autophagy in response to hypoxia, depending on the context [108]. Furthermore, some HIF target genes may not influence cancer pathogenesis whatsoever and may represent genes that are only important in other contexts. This is epitomized by VHL loss in the earliest stages of ccRCC formation, which causes HIF activation in an inappropriate context (i.e. causing a cellular response to hypoxia when the cell is not hypoxic). In this setting, HIF causes a change in cell state that is unwarranted since the cell is exposed to normal oxygen levels. Therefore, many activated HIF target genes may confer no survival advantage or may even result in a “fitness penalty” to the cell in this context. Taken together, the overall consequences of massive HIF pathway activation in ccRCC will be a balance of many positive, neutral and negative effects [109]. The contribution of each effect may change during cancer pathogenesis as a result of subsequent somatic mutation, epigenetic events or changes in the tumor microenvironment allowing cancer cells to escape the long prodromal dormancy that occurs following VHL inactivation. Alternatively, the poise of the HIF transcriptional pathway may be partially pre-set prior to VHL inactivation due to cell-type specific differences in HIF target genes. In turn, this could render specific cell types particularly susceptible to VHL inactivation. Furthermore, genetic differences between individuals might alter specific HIF target genes, thus making that individual more or less susceptible to developing kidney cancer.

Activation of contrasting and aberrant pathways as part of large transcriptional programs is an emerging theme in cancer biology. For example, MYC, like HIF, has transcriptional targets with both oncogenic and tumor suppressive properties [110, 111]. Therefore, HIF activation in ccRCC serves as a model for studying large transcriptional cascades in cancer more generally.

Advertisement

8. Modulation of the HIF response during the pathogenesis of ccRCC

Early evidence to support the pleiotropic nature of the HIF pathway in kidney cancer came from the observation that HIF-1α and HIF-2α have opposing actions on tumor growth in ccRCC xenograft models. Whilst HIF-2α promotes tumor growth, HIF-1α has the opposite effect and restricts tumor growth [46, 48, 50, 51]. Furthermore, expression of HIF-2α target genes in ccRCC tumors correlates with poor patient prognosis, whereas HIF-1α targets genes are associated with improved survival [91].

Commensurate with this, HIF isoform expression appears to switch from HIF-1α to HIF-2α during the development of kidney cancer [28, 112]. In renal tubule epithelial cells, including proximal tubular cells from which ccRCC is derived, HIF-1α mRNA is highly expressed, whereas HIF-2α mRNA is undetectable [28]. Conversely, HIF-2α mRNA (and protein) is highly expressed in ccRCC, possibly as a result of downregulation of DNMT3a and resultant promoter demethylation of the EPAS1 gene that encodes HIF-2α [113]. Furthermore, ccRCCs often downregulate HIF-1α through loss of copy number, deletion, truncation or transcript downregulation [25, 31, 32, 112, 114]. Given the tumor-suppressive function of HIF-1 and the oncogenic function of HIF-2, the shift from HIF-1α in the ccRCC cell of origin to dominant HIF-2α expression in overt ccRCC would favor a more oncogenic phenotype. However, even within the transcriptional repertoire of each isoform there are genes with heterogenous associations with prognosis, suggesting that other selective pressures, effective at the level of individual HIF target genes, may also be operating [91].

Indeed, suppression of individual HIF target genes with anti-tumorigenic properties has been reported in ccRCC. The pro-apoptotic gene BNIP3 is a canonical HIF target gene in many cell types. However, rather than being increased by constitutive HIF in ccRCC cells, its expression was found to be lower than in normal kidney cells. This is most likely as a result of epigenetic modification of the BNIP3 gene locus involving histone deacetylation [115].

In this respect, it is notable that of the many somatic mutations that co-occur with VHL inactivation in ccRCC, very few occur within HIF-target genes. However, to date, the majority of ccRCC sequencing efforts have focused on the coding genome or have targeted genomic regions of interest. Therefore, the majority of HIF binding sites (which are usually intergenic) have not been extensively examined and further studies may reveal somatic mutation of these sites in the future. However, epigenetic modifiers such as PBRM1, SETD2 and BAP1 are recurrently mutated in these tumors [25, 31, 32, 116, 117, 118, 119, 120]. PBRM1 encodes a subunit of the chromatin remodeling PBAF SWI/SNF complex; SETD2 encodes a histone methyltransferase; and BAP1 encodes a histone deubiquitinase. Interestingly, parallel evolution has been reported with respect to these mutations, whereby multiple mutations in the same gene are present in different cells of the same tumor [32]. This emphasizes their importance in driving ccRCC, as well as illustrating their temporal occurrence (i.e. subsequent to VHL mutation). Although the interaction between these ccRCC-associated somatic mutations and the HIF pathway remains unclear, PBRM1 inactivation enhances some aspects of the HIF response [121] and reduces the tumor-suppressor activity of HIF-1α, although the mechanisms are unknown [122]. Recurrent mutations are also found in genes within the PI3K/AKT/mTOR pathway, which is a master regulator of RNA translation. Expression of both HIF-1α and HIF-2α protein are differentially dependent on mTOR, with HIF-1α being regulated by both the mTORC1 and mTORC2 complexes, whilst HIF-2α is dependent solely on mTORC2 [123]. Therefore, HIF isoforms may be differentially affected by mutations in this pathway.

In addition, other oncogenic transcription factors activated in ccRCC may modulate the HIF response. For example, MYC activity is enhanced in ccRCC [124, 125] and synergizes preferentially with HIF-2, whilst antagonizing HIF-1 [102, 126]. In this way, MYC augments the switch from HIF-1 to the more oncogenic HIF-2 isoform. Importantly, MYC itself is a transcriptional target of HIF in ccRCC cells [127], providing a mechanism whereby stabilization of HIF following inactivation of VHL preferentially amplifies the HIF-2 transcriptional pathway in these cells.

Advertisement

9. Variation in the HIF pathway pre-disposes to renal cancer

As discussed above, genetic and epigenetic events occur somatically in ccRCC following VHL inactivation, allowing the HIF transcriptional output to adapt to a more oncogenic phenotype, thereby promoting tumor formation. However, differences in the HIF pathway that exist prior to VHL inactivation can also affect the ability of cells to form cancer. Indeed, it is highly likely that cell-type differences in the HIF pathway contribute to the tight tissue-specificity of VHL-associated cancer, despite the almost universal operation of the VHL-HIF pathway in different mammalian cell types. Potentially, cell-type-specific components of the HIF pathway might favor tumorigenesis in permissive cell types, inhibit tumorigenesis in non-permissive cell types, or a combination of both (Figure 4). The exact mechanism underlying this tissue specificity remains to be determined, although elucidation of HIF target genes in cells permissive to VHL-associated cancers (compared to that in non-permissive cells) will be key in future studies. Of note, the G1/S-phase cell-cycle regulator cyclin D1 (CCND1) has been found to be a HIF-2 responsive gene, which is not regulated by HIF-1 and is unique to ccRCC cells [50]. Furthermore, CCND1 is required for ccRCC cell growth in mice [128]. This indicates that CCND1 and likely other tissue-specific HIF target genes may render certain cell types receptive to tumorigenesis upon VHL inactivation.

Figure 4.

Rebalancing the HIF pathway to favor tumorigenesis. HIF target genes include those that promote tumor growth (depicted in red), restrict tumor growth (depicted in green) and those that do not influence tumor growth (depicted in gray). Depending on the context (i.e. in a permissive or non-permissive context), activation of the HIF pathway may or may not be conducive to tumorigenesis. Features that could ‘tip the balance’ in a HIF-activated cell include genetic mutations, epigenetic features and the cell state (e.g. the underlying gene expression program).

As well as being affected by somatic alterations and cell-type-specific features, the HIF pathway can also be modified by inherited genetic variants. Polymorphisms that predispose individuals to kidney cancer have been studied, and several of these have been shown to affect HIF target genes. Such variants have been identified by genome wide-association studies (GWAS), which compare the genome sequence of renal cancer patients with healthy control individuals [129, 130, 131, 132, 133, 134, 135]. Although these variants likely only account for about 5% of kidney cancer heritability [129], a disproportionately high number of these susceptibility loci overlap with cis-acting components of the HIF pathway [136]. This indicates that specific aspects of the HIF pathway are under genetic selection during the development of kidney cancer.

Many of these RCC-susceptibility loci lie in intergenic regions and so the functional target of these polymorphisms is not immediately apparent. However, several susceptibility loci overlap with, or lie adjacent to, HIF-binding sites [136]. In-depth analysis of chromatin looping and HIF-dependent gene regulation has identified a number of HIF target genes associated with these loci [127, 136, 137, 138]. At each locus, the renal cancer susceptibility polymorphism affects both HIF binding and expression of the HIF-target gene, either by generating a second HRE motif or by altering chromatin accessibility. Most notable are polymorphisms at the 11q13.3 locus, which affect HIF-2-dependent expression of cyclin D1 (CCND1) [137]; polymorphisms at the 8q24.21 locus, which affect HIF binding and expression of the oncogenic transcription factor MYC [127]; and polymorphisms at the 12p12.1 locus, which alter HIF-1 dependent expression of the basic helix–loop–helix transcription factor BHLHE41 (also known as DEC2) [138]. Furthermore, RCC-susceptibility polymorphisms have been identified at the 2p21 locus, lying in the first intron of the EPAS1 gene that encodes HIF-2α, although whether these affect HIF-2α expression remains unclear [130, 139]. Importantly, each renal cancer susceptibility locus affects a single component of the HIF pathway. This directly implicates these genes in the pathogenesis of kidney cancer. Furthermore, it helps distinguish them from HIF target genes with neutral effects on RCC susceptibility that might be simply co-activated as part of large pathway upregulation. Therefore, these analyses have highlighted specific ‘driver’ genes that may provide attractive targets for future therapeutic approaches or as biomarkers that might predict tumor behavior.

Advertisement

10. Therapeutic implications of HIF pathway activation in ccRCC

In the absence of a surgical cure, the outlook for patients with clear cell renal cancer is poor, with a median survival of just 2 years. However, over recent years a number of systemic anti-cancer therapeutic strategies have emerged, which are beginning to alter the outcome for some of these patients.

10.1 Anti-angiogenic therapies

One strategy has focused on angiogenesis inhibitors to treat metastatic ccRCC. Whilst all tumors require a blood supply to obtain sufficient oxygen and nutrients to grow, ccRCC (and other VHL-dependent cancers such as hemangioblastoma) are particularly rich in blood vessels. Indeed, VEGFA, a master regulator of angiogenesis, [98, 99] is a direct transcriptional target of HIF and is highly expressed in ccRCC cells [41, 140]. Early anti-angiogenic strategies targeted VEGFA using the monoclonal antibody bevacizumab, with limited efficacy [141]. However, several other HIF target genes also encode pro-angiogenic factors, such as PGF, adrenomedullin (ADM) and plasminogen activator 1 (PAI-1), as well as the VEGF receptor, FLT1. These likely act in concert with VEGFA to orchestrate a robust angiogenic phenotype in the context of HIF activation. Therefore, rather than targeting individual factors, more recent strategies have used small-molecule receptor tyrosine kinase inhibitors (TKIs) to block the overarching angiogenic pathways [142]. However, while effective in some individuals, other tumors may fail to respond, likely reflecting heterogeneity in gene expression between tumors. Furthermore, the duration of response may be limited, possibly reflecting intra-tumor heterogeneity and the growth of resistant subclones.

10.2 Immunotherapy

In recent years, immune checkpoint inhibition via targeting PD-L1 and CTLA-4 has emerged as an effective treatment for advanced ccRCC. This is despite the relatively low mutational burden seen in this type of cancer, which often correlates with sensitivity to immunomodulatory therapy in other cancer types. Whilst HIF has multiple effects on the immune response [143], it is of particular interest that PD-L1 has been found to be transcriptionally regulated by HIF in ccRCC cells [107, 144, 145]. Therefore, it is possible that HIF-mediated activation of PD-L1 may underlie the sensitivity of ccRCC to inhibition of this pathway.

10.3 mTOR inhibitors

Historically, mTOR inhibitors have been used in the treatment of metastatic kidney cancer and remain part of the modern armamentarium [146, 147]. Inhibition of mTOR will negatively impact translation of HIF-alpha subunits, while preferential blockade of mTORC1 or mTORC2 may alter the balance of the two isoforms. Given the oncogenic role of HIF-2α in ccRCC and the selective regulation of HIF-2α by mTORC2, mTORC2 inhibition may provide a more targeted therapeutic approach in the future.

10.4 HIF-2 inhibitors

The finding that HIF-1α and HIF-2α have opposing effects on the pathogenesis of ccRCC initiated efforts to generate isoform-specific inhibitors. This led to the development of small molecule inhibitors that specifically prevent HIF-2α dimerizing with HIF-1β, thereby blocking HIF-2α -dependent transcription without affecting HIF-1α activity [148]. These inhibitors would be predicted to have greater efficacy compared to targeting both isoforms simultaneously, whilst reducing off-target side-effects. Indeed, investigation of these compounds as potential ccRCC treatments, both in animal models of ccRCC and early clinical trials, have yielded promising results [149, 150, 151]. Therefore, these compounds could provide another strategy for treating metastatic ccRCC.

11. Conclusions

Inactivation of the VHL tumor suppressor gene is the hallmark of clear cell renal cancer and leads to the upregulation of wide-spread hypoxia pathways, orchestrated by the transcription factor HIF. Whilst HIF proteins activate many genes that are central to the “hallmarks of cancer”, other HIF-target genes may restrict cancer progression and the overall consequence of HIF pathway activation is a balance of these effects (Figure 4). Both genetic and epigenetic genetic events, occurring before or after VHL loss and HIF activation, can alter this balance to promote tumorigenesis.

References

  1. 1. Bray F, Ferlay J, Soerjomataram I, Siegel RL, Torre LA, Jemal A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2018;68(6):394-424
  2. 2. Drewniak T, Sandheim M, Jakubowski J, Juszczak K, Stelmach AW. Prognostic factors of overall survival in renal cancer patients - single oncological center study. Cent European J Urol. 2013;66(3):283-91
  3. 3. Bedke J, Albiges L, Capitanio U, Giles RH, Hora M, Lam TB, et al. Updated European Association of Urology Guidelines on Renal Cell Carcinoma: Nivolumab plus Cabozantinib Joins Immune Checkpoint Inhibition Combination Therapies for Treatment-naive Metastatic Clear-Cell Renal Cell Carcinoma. Eur Urol. 2020
  4. 4. Ericsson JL, Seljelid R, Orrenius S. Comparative light and electron microscopic observations of the cytoplasmic matrix in renal carcinomas. Virchows Arch Pathol Anat Physiol Klin Med. 1966;341(3):204-23
  5. 5. Pode-Shakked N, Dekel B. Wilms tumor--a renal stem cell malignancy? Pediatr Nephrol. 2011;26(9):1535-43
  6. 6. Cai Q, Christie A, Rajaram S, Zhou Q, Araj E, Chintalapati S, et al. Ontological analyses reveal clinically-significant clear cell renal cell carcinoma subtypes with convergent evolutionary trajectories into an aggressive type. EBioMedicine. 2020;51:102526
  7. 7. Choyke PL, Glenn GM, Walther MM, Zbar B, Linehan WM. Hereditary renal cancers. Radiology. 2003;226(1):33-46
  8. 8. Melmon KL, Rosen SW. Lindau's Disease. Review of the Literature and Study of a Large Kindred. Am J Med. 1964;36:595-617
  9. 9. Maher ER, Yates JR, Harries R, Benjamin C, Harris R, Moore AT, et al. Clinical features and natural history of von Hippel-Lindau disease. Q J Med. 1990;77(283):1151-63
  10. 10. Glenn GM, Daniel LN, Choyke P, Linehan WM, Oldfield E, Gorin MB, et al. Von Hippel-Lindau (VHL) disease: distinct phenotypes suggest more than one mutant allele at the VHL locus. Human genetics. 1991;87(2):207-10
  11. 11. Neumann HP, Wiestler OD. Clustering of features of von Hippel-Lindau syndrome: evidence for a complex genetic locus. Lancet. 1991;337(8749):1052-4
  12. 12. Clifford SC, Cockman ME, Smallwood AC, Mole DR, Woodward ER, Maxwell PH, et al. Contrasting effects on HIF-1alpha regulation by disease-causing pVHL mutations correlate with patterns of tumourigenesis in von Hippel-Lindau disease. Hum Mol Genet. 2001;10(10):1029-38
  13. 13. Chen F, Kishida T, Yao M, Hustad T, Glavac D, Dean M, et al. Germline mutations in the von Hippel-Lindau disease tumor suppressor gene: correlations with phenotype. Hum Mutat. 1995;5(1):66-75
  14. 14. Zbar B, Kishida T, Chen F, Schmidt L, Maher ER, Richards FM, et al. Germline mutations in the Von Hippel-Lindau disease (VHL) gene in families from North America, Europe, and Japan. Hum Mutat. 1996;8(4):348-57
  15. 15. Neumann HP, Bender BU. Genotype-phenotype correlations in von Hippel-Lindau disease. J Intern Med. 1998;243(6):541-5
  16. 16. Hoffman MA, Ohh M, Yang H, Klco JM, Ivan M, Kaelin WG, Jr. von Hippel-Lindau protein mutants linked to type 2C VHL disease preserve the ability to downregulate HIF. Hum Mol Genet. 2001;10(10):1019-27
  17. 17. Li L, Zhang L, Zhang X, Yan Q, Minamishima YA, Olumi AF, et al. Hypoxia-inducible factor linked to differential kidney cancer risk seen with type 2A and type 2B VHL mutations. Molecular and cellular biology. 2007;27(15):5381-92
  18. 18. Latif F, Tory K, Gnarra J, Yao M, Duh FM, Orcutt ML, et al. Identification of the von Hippel-Lindau disease tumor suppressor gene. Science. 1993;260(5112):1317-20
  19. 19. Schoenfeld A, Davidowitz EJ, Burk RD. A second major native von Hippel-Lindau gene product, initiated from an internal translation start site, functions as a tumor suppressor. Proceedings of the National Academy of Sciences of the United States of America. 1998;95(15):8817-22
  20. 20. Maher ER, Yates JR, Ferguson-Smith MA. Statistical analysis of the two stage mutation model in von Hippel-Lindau disease, and in sporadic cerebellar haemangioblastoma and renal cell carcinoma. J Med Genet. 1990;27(5):311-4
  21. 21. Tory K, Brauch H, Linehan M, Barba D, Oldfield E, Filling-Katz M, et al. Specific genetic change in tumors associated with von Hippel-Lindau disease. J Natl Cancer Inst. 1989;81(14):1097-101
  22. 22. Knudson AG, Jr. Mutation and cancer: statistical study of retinoblastoma. Proceedings of the National Academy of Sciences of the United States of America. 1971;68(4):820-3
  23. 23. Knudson AG, Jr. Genetics of human cancer. Annual review of genetics. 1986;20:231-51
  24. 24. Young MD, Mitchell TJ, Vieira Braga FA, Tran MGB, Stewart BJ, Ferdinand JR, et al. Single-cell transcriptomes from human kidneys reveal the cellular identity of renal tumors. Science. 2018;361(6402):594-9
  25. 25. Sato Y, Yoshizato T, Shiraishi Y, Maekawa S, Okuno Y, Kamura T, et al. Integrated molecular analysis of clear-cell renal cell carcinoma. Nature genetics. 2013;45(8):860-7
  26. 26. Mitchell TJ, Turajlic S, Rowan A, Nicol D, Farmery JHR, O'Brien T, et al. Timing the Landmark Events in the Evolution of Clear Cell Renal Cell Cancer: TRACERx Renal. Cell. 2018;173(3):611-23
  27. 27. Mandriota SJ, Turner KJ, Davies DR, Murray PG, Morgan NV, Sowter HM, et al. HIF activation identifies early lesions in VHL kidneys: evidence for site-specific tumor suppressor function in the nephron. Cancer cell. 2002;1(5):459-68
  28. 28. Schietke RE, Hackenbeck T, Tran M, Gunther R, Klanke B, Warnecke CL, et al. Renal Tubular HIF-2alpha Expression Requires VHL Inactivation and Causes Fibrosis and Cysts. PLoS ONE. 2012;7(1):e31034
  29. 29. Mack FA, Patel JH, Biju MP, Haase VH, Simon MC. Decreased growth of Vhl−/− fibrosarcomas is associated with elevated levels of cyclin kinase inhibitors p21 and p27. Molecular and cellular biology. 2005;25(11):4565-78
  30. 30. Young AP, Schlisio S, Minamishima YA, Zhang Q, Li L, Grisanzio C, et al. VHL loss actuates a HIF-independent senescence programme mediated by Rb and p400. Nature cell biology. 2008;10(3):361-9
  31. 31. Clark DJ, Dhanasekaran SM, Petralia F, Pan J, Song X, Hu Y, et al. Integrated Proteogenomic Characterization of Clear Cell Renal Cell Carcinoma. Cell. 2019;179(4):964-83
  32. 32. Turajlic S, Xu H, Litchfield K, Rowan A, Horswell S, Chambers T, et al. Deterministic Evolutionary Trajectories Influence Primary Tumor Growth: TRACERx Renal. Cell. 2018;173(3):595-610
  33. 33. Cancer_Genome_Atlas_Research_Network. Comprehensive molecular characterization of clear cell renal cell carcinoma. Nature. 2013;499(7456):43-9
  34. 34. 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(5229):1444-6
  35. 35. Pause A, Lee S, Worrell RA, Chen DY, Burgess WH, Linehan WM, et al. The von Hippel-Lindau tumor-suppressor gene product forms a stable complex with human CUL-2, a member of the Cdc53 family of proteins. Proceedings of the National Academy of Sciences of the United States of America. 1997;94(6):2156-61
  36. 36. Iwai K, Yamanaka K, Kamura T, Minato N, Conaway RC, Conaway JW, et al. Identification of the von Hippel-lindau tumor-suppressor protein as part of an active E3 ubiquitin ligase complex. Proceedings of the National Academy of Sciences of the United States of America. 1999;96(22):12436-41
  37. 37. Kamura T, Koepp DM, Conrad MN, Skowyra D, Moreland RJ, Iliopoulos O, et al. Rbx1, a component of the VHL tumor suppressor complex and SCF ubiquitin ligase. Science. 1999;284(5414):657-61
  38. 38. Hakimi AA, Tickoo SK, Jacobsen A, Sarungbam J, Sfakianos JP, Sato Y, et al. TCEB1-mutated renal cell carcinoma: a distinct genomic and morphological subtype. Mod Pathol. 2015;28(6):845-53
  39. 39. Kamura T, Sato S, Haque D, Liu L, Kaelin WG, Jr., Conaway RC, et al. The Elongin BC complex interacts with the conserved SOCS-box motif present in members of the SOCS, ras, WD-40 repeat, and ankyrin repeat families. Genes & development. 1998;12(24):3872-81
  40. 40. Stebbins CE, Kaelin WG, Jr., Pavletich NP. Structure of the VHL-ElonginC-ElonginB complex: implications for VHL tumor suppressor function. Science. 1999;284(5413):455-61
  41. 41. Maxwell PH, Wiesener MS, Chang GW, Clifford SC, Vaux EC, Cockman ME, et al. The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature. 1999;399(6733):271-5
  42. 42. Cockman ME, Masson N, Mole DR, Jaakkola P, Chang GW, Clifford SC, et al. Hypoxia inducible factor-alpha binding and ubiquitylation by the von Hippel-Lindau tumor suppressor protein. The Journal of biological chemistry. 2000;275(33):25733-41
  43. 43. Ohh M, Park CW, Ivan M, Hoffman MA, Kim TY, Huang LE, et al. Ubiquitination of hypoxia-inducible factor requires direct binding to the beta-domain of the von Hippel-Lindau protein. Nature cell biology. 2000;2(7):423-7
  44. 44. Jaakkola P, Mole DR, Tian YM, Wilson MI, Gielbert J, Gaskell SJ, et al. Targeting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science. 2001;292(5516):468-72
  45. 45. Ivan M, Kondo K, Yang H, Kim W, Valiando J, Ohh M, et al. HIFalpha targeted for VHL-mediated destruction by proline hydroxylation: implications for O2 sensing. Science. 2001;292(5516):464-8
  46. 46. Biswas S, Troy H, Leek R, Chung YL, Li JL, Raval RR, et al. Effects of HIF-1alpha and HIF2alpha on Growth and Metabolism of Clear-Cell Renal Cell Carcinoma 786-0 Xenografts. Journal of oncology. 2010;2010:757908
  47. 47. Kondo K, Klco J, Nakamura E, Lechpammer M, Kaelin WG, Jr. Inhibition of HIF is necessary for tumor suppression by the von Hippel-Lindau protein. Cancer cell. 2002;1(3):237-46
  48. 48. Kondo K, Kim WY, Lechpammer M, Kaelin WG, Jr. Inhibition of HIF2alpha is sufficient to suppress pVHL-defective tumor growth. PLoS biology. 2003;1(3):E83
  49. 49. Zimmer M, Doucette D, Siddiqui N, Iliopoulos O. Inhibition of hypoxia-inducible factor is sufficient for growth suppression of VHL−/− tumors. Mol Cancer Res. 2004;2(2):89-95
  50. 50. Raval RR, Lau KW, Tran MG, Sowter HM, Mandriota SJ, Li JL, et al. Contrasting Properties of Hypoxia-Inducible Factor 1 (HIF-1) and HIF-2 in von Hippel-Lindau-Associated Renal Cell Carcinoma. Molecular and cellular biology. 2005;25(13):5675-86
  51. 51. Hoefflin R, Harlander S, Schafer S, Metzger P, Kuo F, Schonenberger D, et al. HIF-1alpha and HIF-2alpha differently regulate tumour development and inflammation of clear cell renal cell carcinoma in mice. Nature communications. 2020;11(1):4111
  52. 52. Nyhan MJ, O'Sullivan GC, McKenna SL. Role of the VHL (von Hippel-Lindau) gene in renal cancer: a multifunctional tumour suppressor. Biochem Soc Trans. 2008;36(Pt 3):472-8
  53. 53. Li Z, Wang D, Na X, Schoen SR, Messing EM, Wu G. Identification of a deubiquitinating enzyme subfamily as substrates of the von Hippel-Lindau tumor suppressor. Biochemical and biophysical research communications. 2002;294(3):700-9
  54. 54. Li Z, Na X, Wang D, Schoen SR, Messing EM, Wu G. Ubiquitination of a novel deubiquitinating enzyme requires direct binding to von Hippel-Lindau tumor suppressor protein. The Journal of biological chemistry. 2002;277(7):4656-62
  55. 55. Li Z, Wang D, Messing EM, Wu G. VHL protein-interacting deubiquitinating enzyme 2 deubiquitinates and stabilizes HIF-1alpha. EMBO Rep. 2005;6(4):373-8
  56. 56. Na X, Duan HO, Messing EM, Schoen SR, Ryan CK, di Sant'Agnese PA, et al. Identification of the RNA polymerase II subunit hsRPB7 as a novel target of the von Hippel-Lindau protein. The EMBO journal. 2003;22(16):4249-59
  57. 57. Kuznetsova AV, Meller J, Schnell PO, Nash JA, Ignacak ML, Sanchez Y, et al. von Hippel-Lindau protein binds hyperphosphorylated large subunit of RNA polymerase II through a proline hydroxylation motif and targets it for ubiquitination. Proceedings of the National Academy of Sciences of the United States of America. 2003;100(5):2706-11
  58. 58. Mikhaylova O, Ignacak ML, Barankiewicz TJ, Harbaugh SV, Yi Y, Maxwell PH, et al. The von Hippel-Lindau tumor suppressor protein and Egl-9-Type proline hydroxylases regulate the large subunit of RNA polymerase II in response to oxidative stress. Molecular and cellular biology. 2008;28(8):2701-17
  59. 59. Heir P, Srikumar T, Bikopoulos G, Bunda S, Poon BP, Lee JE, et al. Oxygen-dependent Regulation of Erythropoietin Receptor Turnover and Signaling. The Journal of biological chemistry. 2016;291(14):7357-72
  60. 60. Okuda H, Hirai S, Takaki Y, Kamada M, Baba M, Sakai N, et al. Direct interaction of the beta-domain of VHL tumor suppressor protein with the regulatory domain of atypical PKC isotypes. Biochemical and biophysical research communications. 1999;263(2):491-7
  61. 61. Okuda H, Saitoh K, Hirai S, Iwai K, Takaki Y, Baba M, et al. The von Hippel-Lindau tumor suppressor protein mediates ubiquitination of activated atypical protein kinase C. The Journal of biological chemistry. 2001;276(47):43611-7
  62. 62. Guo J, Chakraborty AA, Liu P, Gan W, Zheng X, Inuzuka H, et al. pVHL suppresses kinase activity of Akt in a proline-hydroxylation-dependent manner. Science. 2016;353(6302):929-32
  63. 63. Anderson K, Nordquist KA, Gao X, Hicks KC, Zhai B, Gygi SP, et al. Regulation of cellular levels of Sprouty2 protein by prolyl hydroxylase domain and von Hippel-Lindau proteins. The Journal of biological chemistry. 2011;286(49):42027-36
  64. 64. Kurban G, Duplan E, Ramlal N, Hudon V, Sado Y, Ninomiya Y, et al. Collagen matrix assembly is driven by the interaction of von Hippel-Lindau tumor suppressor protein with hydroxylated collagen IV alpha 2. Oncogene. 2008;27(7):1004-12
  65. 65. Grosfeld A, Stolze IP, Cockman ME, Pugh CW, Edelmann M, Kessler B, et al. Interaction of hydroxylated collagen IV with the von hippel-lindau tumor suppressor. The Journal of biological chemistry. 2007;282(18):13264-9
  66. 66. Ohh M, Yauch RL, Lonergan KM, Whaley JM, Stemmer-Rachamimov AO, Louis DN, et al. The von Hippel-Lindau tumor suppressor protein is required for proper assembly of an extracellular fibronectin matrix. Molecular cell. 1998;1(7):959-68
  67. 67. 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. Proceedings of the National Academy of Sciences of the United States of America. 1995;92(12):5510-4
  68. 68. Wiesener MS, Turley H, Allen WE, Willam C, Eckardt KU, Talks KL, et al. Induction of endothelial PAS domain protein-1 by hypoxia: characterization and comparison with hypoxia-inducible factor-1alpha. Blood. 1998;92(7):2260-8
  69. 69. Tian H, McKnight SL, Russell DW. Endothelial PAS domain protein 1 (EPAS1), a transcription factor selectively expressed in endothelial cells. Genes & development. 1997;11(1):72-82
  70. 70. Maynard MA, Qi H, Chung J, Lee EH, Kondo Y, Hara S, et al. Multiple splice variants of the human HIF-3 alpha locus are targets of the von Hippel-Lindau E3 ubiquitin ligase complex. The Journal of biological chemistry. 2003;278(13):11032-40
  71. 71. Benita Y, Kikuchi H, Smith AD, Zhang MQ, Chung DC, Xavier RJ. An integrative genomics approach identifies Hypoxia Inducible Factor-1 (HIF-1)-target genes that form the core response to hypoxia. Nucleic acids research. 2009;37(14):4587-602
  72. 72. 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 expression. 1998;7(3):205-13
  73. 73. Maynard MA, Evans AJ, Hosomi T, Hara S, Jewett MA, Ohh M. Human HIF-3alpha4 is a dominant-negative regulator of HIF-1 and is down-regulated in renal cell carcinoma. Faseb J. 2005;19(11):1396-406
  74. 74. Makino Y, Cao R, Svensson K, Bertilsson G, Asman M, Tanaka H, et al. Inhibitory PAS domain protein is a negative regulator of hypoxia-inducible gene expression. Nature. 2001;414(6863):550-4
  75. 75. Heikkila M, Pasanen A, Kivirikko KI, Myllyharju J. Roles of the human hypoxia-inducible factor (HIF)-3alpha variants in the hypoxia response. Cell Mol Life Sci. 2011;68(23):3885-901
  76. 76. Bersten DC, Sullivan AE, Peet DJ, Whitelaw ML. bHLH-PAS proteins in cancer. Nature reviews. 2013;13(12):827-41
  77. 77. Kaelin WG, Jr., Ratcliffe PJ. Oxygen sensing by metazoans: the central role of the HIF hydroxylase pathway. Molecular cell. 2008;30(4):393-402
  78. 78. Duan C. Hypoxia-inducible factor 3 biology: complexities and emerging themes. Am J Physiol Cell Physiol. 2016;310(4):C260-9
  79. 79. Epstein AC, Gleadle JM, McNeill LA, Hewitson KS, O'Rourke J, Mole DR, et al. C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell. 2001;107(1):43-54
  80. 80. 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 & development. 2002;16(12):1466-71
  81. 81. Lando D, Peet DJ, Whelan DA, Gorman JJ, Whitelaw ML. Asparagine hydroxylation of the HIF transactivation domain a hypoxic switch. Science. 2002;295(5556):858-61
  82. 82. 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 & development. 2001;15(20):2675-86
  83. 83. Khan MN, Bhattacharyya T, Andrikopoulos P, Esteban MA, Barod R, Connor T, et al. Factor inhibiting HIF (FIH-1) promotes renal cancer cell survival by protecting cells from HIF-1alpha-mediated apoptosis. British journal of cancer. 2011;104(7):1151-9
  84. 84. Wenger RH, Stiehl DP, Camenisch G. Integration of oxygen signaling at the consensus HRE. Sci STKE. 2005;2005(306):re12
  85. 85. Schodel J, Oikonomopoulos S, Ragoussis J, Pugh CW, Ratcliffe PJ, Mole DR. High-resolution genome-wide mapping of HIF-binding sites by ChIP-seq. Blood. 2010;117(23):e207-17
  86. 86. Choudhry H, Schodel J, Oikonomopoulos S, Camps C, Grampp S, Harris AL, et al. Extensive regulation of the non-coding transcriptome by hypoxia: role of HIF in releasing paused RNApol2. EMBO Rep. 2014;15(1):70-6
  87. 87. Xia X, Kung AL. Preferential binding of HIF-1 to transcriptionally active loci determines cell-type specific response to hypoxia. Genome biology. 2009;10(10):R113
  88. 88. Xia X, Lemieux ME, Li W, Carroll JS, Brown M, Liu XS, et al. Integrative analysis of HIF binding and transactivation reveals its role in maintaining histone methylation homeostasis. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(11):4260-5
  89. 89. Platt JL, Salama R, Smythies J, Choudhry H, Davies JO, Hughes JR, et al. Capture-C reveals preformed chromatin interactions between HIF-binding sites and distant promoters. EMBO Rep. 2016;17(10):1410-21
  90. 90. Smythies JA, Sun M, Masson N, Salama R, Simpson PD, Murray E, et al. Inherent DNA-binding specificities of the HIF-1alpha and HIF-2alpha transcription factors in chromatin. EMBO Rep. 2019;20(1):e46401
  91. 91. Salama R, Masson N, Simpson P, Sciesielski LK, Sun M, Tian YM, et al. Heterogeneous Effects of Direct Hypoxia Pathway Activation in Kidney Cancer. PLoS ONE. 2015;10(8):e0134645
  92. 92. Schodel J, Mole DR, Ratcliffe PJ. Pan-genomic binding of hypoxia-inducible transcription factors. Biol Chem. 2013;394(4):507-17
  93. 93. Ortiz-Barahona A, Villar D, Pescador N, Amigo J, del Peso L. Genome-wide identification of hypoxia-inducible factor binding sites and target genes by a probabilistic model integrating transcription-profiling data and in silico binding site prediction. Nucleic acids research. 2010;38(7):2332-45
  94. 94. Schofield CJ, Ratcliffe PJ. Oxygen sensing by HIF hydroxylases. Nat Rev Mol Cell Biol. 2004;5(5):343-54
  95. 95. Semenza GL. Hypoxia-inducible factors in physiology and medicine. Cell. 2012;148(3):399-408
  96. 96. Lau KW, Tian YM, Raval RR, Ratcliffe PJ, Pugh CW. Target gene selectivity of hypoxia-inducible factor-alpha in renal cancer cells is conveyed by post-DNA-binding mechanisms. British journal of cancer. 2007;96(8):1284-92
  97. 97. Sowter HM, Raval RR, Moore JW, Ratcliffe PJ, Harris AL. Predominant role of hypoxia-inducible transcription factor (Hif)-1alpha versus Hif-2alpha in regulation of the transcriptional response to hypoxia. Cancer research. 2003;63(19):6130-4
  98. 98. Gleadle JM, Ebert BL, Firth JD, Ratcliffe PJ. Regulation of angiogenic growth factor expression by hypoxia, transition metals, and chelating agents. The American journal of physiology. 1995;268(6 Pt 1):C1362-8
  99. 99. Forsythe JA, Jiang BH, Iyer NV, Agani F, Leung SW, Koos RD, et al. Activation of vascular endothelial growth factor gene transcription by hypoxia-inducible factor 1. Molecular and cellular biology. 1996;16(9):4604-13
  100. 100. Sowter HM, Ratcliffe PJ, Watson P, Greenberg AH, Harris AL. HIF-1-dependent regulation of hypoxic induction of the cell death factors BNIP3 and NIX in human tumors. Cancer research. 2001;61(18):6669-73
  101. 101. Covello KL, Kehler J, Yu H, Gordan JD, Arsham AM, Hu CJ, et al. HIF-2alpha regulates Oct-4: effects of hypoxia on stem cell function, embryonic development, and tumor growth. Genes & development. 2006;20(5):557-70
  102. 102. 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(4):335-47
  103. 103. Semenza GL. Hypoxia-inducible factors: mediators of cancer progression and targets for cancer therapy. Trends Pharmacol Sci. 2012;33(4):207-14
  104. 104. Semenza GL. Defining the role of hypoxia-inducible factor 1 in cancer biology and therapeutics. Oncogene. 2010;29:625-34
  105. 105. Chiche J, Ilc K, Laferriere J, Trottier E, Dayan F, Mazure NM, et al. Hypoxia-inducible carbonic anhydrase IX and XII promote tumor cell growth by counteracting acidosis through the regulation of the intracellular pH. Cancer research. 2009;69(1):358-68
  106. 106. Krock BL, Skuli N, Simon MC. Hypoxia-induced angiogenesis: good and evil. Genes & cancer. 2011;2(12):1117-33
  107. 107. Ruf M, Moch H, Schraml P. PD-L1 expression is regulated by hypoxia inducible factor in clear cell renal cell carcinoma. International journal of cancer. 2016;139(2):396-403
  108. 108. Zhang J, Ney PA. Role of BNIP3 and NIX in cell death, autophagy, and mitophagy. Cell death and differentiation. 2009;16(7):939-46
  109. 109. Maxwell PH, Pugh CW, Ratcliffe PJ. Activation of the HIF pathway in cancer. Curr Opin Genet Dev. 2001;11(3):293-9
  110. 110. Murphy DJ, Junttila MR, Pouyet L, Karnezis A, Shchors K, Bui DA, et al. Distinct thresholds govern Myc's biological output in vivo. Cancer cell. 2008;14(6):447-57
  111. 111. Dang CV. c-Myc target genes involved in cell growth, apoptosis, and metabolism. Molecular and cellular biology. 1999;19(1):1-11
  112. 112. Shen C, Beroukhim R, Schumacher SE, Zhou J, Chang M, Signoretti S, et al. Genetic and Functional Studies Implicate HIF1alpha as a 14q Kidney Cancer Suppressor Gene. Cancer Discov. 2011;1(3):222-35
  113. 113. Lachance G, Uniacke J, Audas TE, Holterman CE, Franovic A, Payette J, et al. DNMT3a epigenetic program regulates the HIF-2alpha oxygen-sensing pathway and the cellular response to hypoxia. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(21):7783-8
  114. 114. Shinojima T, Oya M, Takayanagi A, Mizuno R, Shimizu N, Murai M. Renal cancer cells lacking hypoxia inducible factor (HIF)-1alpha expression maintain vascular endothelial growth factor expression through HIF-2alpha. Carcinogenesis. 2007;28(3):529-36
  115. 115. Shao Y, Liu Z, Liu J, Wang H, Huang L, Lin T, et al. Expression and epigenetic regulatory mechanism of BNIP3 in clear cell renal cell carcinoma. Int J Oncol. 2019;54(1):348-60
  116. 116. COSMIC. Catalogue Of Somatic Mutations In Cancer 2012 [Available from: http://www.sanger.ac.uk/genetics/CGP/cosmic/
  117. 117. Varela I, Tarpey P, Raine K, Huang D, Ong CK, Stephens P, et al. Exome sequencing identifies frequent mutation of the SWI/SNF complex gene PBRM1 in renal carcinoma. Nature. 2011;469(7331):539-42
  118. 118. Duns G, Hofstra RM, Sietzema JG, Hollema H, van Duivenbode I, Kuik A, et al. Targeted exome sequencing in clear cell renal cell carcinoma tumors suggests aberrant chromatin regulation as a crucial step in ccRCC development. Hum Mutat. 2012;33(7):1059-62
  119. 119. Pena-Llopis S, Vega-Rubin-de-Celis S, Liao A, Leng N, Pavia-Jimenez A, Wang S, et al. BAP1 loss defines a new class of renal cell carcinoma. Nature genetics. 2012;44(7):751-9
  120. 120. Dalgliesh GL, Furge K, Greenman C, Chen L, Bignell G, Butler A, et al. Systematic sequencing of renal carcinoma reveals inactivation of histone modifying genes. Nature. 2010;463(7279):360-3
  121. 121. Gao W, Li W, Xiao T, Liu XS, Kaelin WG, Jr. Inactivation of the PBRM1 tumor suppressor gene amplifies the HIF-response in VHL−/− clear cell renal carcinoma. Proceedings of the National Academy of Sciences of the United States of America. 2017;114(5):1027-32
  122. 122. Murakami A, Wang L, Kalhorn S, Schraml P, Rathmell WK, Tan AC, et al. Context-dependent role for chromatin remodeling component PBRM1/BAF180 in clear cell renal cell carcinoma. Oncogenesis. 2017;6(1):e287
  123. 123. Toschi A, Lee E, Gadir N, Ohh M, Foster DA. Differential dependence of hypoxia-inducible factors 1 alpha and 2 alpha on mTORC1 and mTORC2. The Journal of biological chemistry. 2008;283(50):34495-9
  124. 124. Bailey ST, Smith AM, Kardos J, Wobker SE, Wilson HL, Krishnan B, et al. MYC activation cooperates with Vhl and Ink4a/Arf loss to induce clear cell renal cell carcinoma. Nature communications. 2017;8:15770
  125. 125. Tang SW, Chang WH, Su YC, Chen YC, Lai YH, Wu PT, et al. MYC pathway is activated in clear cell renal cell carcinoma and essential for proliferation of clear cell renal cell carcinoma cells. Cancer Lett. 2009;273(1):35-43
  126. 126. Dang CV, Kim JW, Gao P, Yustein J. The interplay between MYC and HIF in cancer. Nature reviews. 2008;8(1):51-6
  127. 127. Grampp S, Platt JL, Lauer V, Salama R, Kranz F, Neumann VK, et al. Genetic variation at the 8q24.21 renal cancer susceptibility locus affects HIF binding to a MYC enhancer. Nature communications. 2016;7:13183
  128. 128. Zhang T, Niu X, Liao L, Cho EA, Yang H. The contributions of HIF-target genes to tumor growth in RCC. PLoS ONE. 2013;8(11):e80544
  129. 129. Scelo G, Purdue MP, Brown KM, Johansson M, Wang Z, Eckel-Passow JE, et al. Genome-wide association study identifies multiple risk loci for renal cell carcinoma. Nature communications. 2017;8:15724
  130. 130. Purdue MP, Johansson M, Zelenika D, Toro JR, Scelo G, Moore LE, et al. Genome-wide association study of renal cell carcinoma identifies two susceptibility loci on 2p21 and 11q13.3. Nature genetics. 2011;43(1):60-5
  131. 131. Wu X, Scelo G, Purdue MP, Rothman N, Johansson M, Ye Y, et al. A genome-wide association study identifies a novel susceptibility locus for renal cell carcinoma on 12p11.23. Hum Mol Genet. 2011;21(2):456-62
  132. 132. Henrion M, Frampton M, Scelo G, Purdue M, Ye Y, Broderick P, et al. Common variation at 2q22.3 (ZEB2) influences the risk of renal cancer. Hum Mol Genet. 2013;22(4):825-31
  133. 133. Henrion MY, Purdue MP, Scelo G, Broderick P, Frampton M, Ritchie A, et al. Common variation at 1q24.1 (ALDH9A1) is a potential risk factor for renal cancer. PLoS ONE. 2015;10(3):e0122589
  134. 134. Gudmundsson J, Sulem P, Gudbjartsson DF, Blondal T, Gylfason A, Agnarsson BA, et al. Genome-wide association and replication studies identify four variants associated with prostate cancer susceptibility. Nature genetics. 2009;41(10):1122-6
  135. 135. Gudmundsson J, Sulem P, Gudbjartsson DF, Masson G, Petursdottir V, Hardarson S, et al. A common variant at 8q24.21 is associated with renal cell cancer. Nature communications. 2013;4:2776
  136. 136. Schmid V, Lafleur VN, Lombardi O, Li R, Salama R, Colli L, et al. Co-incidence of RCC-susceptibility polymorphisms with HIF cis-acting sequences supports a pathway tuning model of cancer. Scientific reports. 2019;9(1):18768
  137. 137. Schodel J, Bardella C, Sciesielski LK, Brown JM, Pugh CW, Buckle V, et al. Common genetic variants at the 11q13.3 renal cancer susceptibility locus influence binding of HIF to an enhancer of cyclin D1 expression. Nature genetics. 2012;44(4):420-5
  138. 138. Grampp S, Schmid V, Salama R, Lauer V, Kranz F, Platt JL, et al. Multiple renal cancer susceptibility polymorphisms modulate the HIF pathway. PLoS Genet. 2017;13(7):e1006872
  139. 139. Han SS, Yeager M, Moore LE, Wei MH, Pfeiffer R, Toure O, et al. The chromosome 2p21 region harbors a complex genetic architecture for association with risk for renal cell carcinoma. Hum Mol Genet. 2011;21(5):1190-200
  140. 140. Iliopoulos O, Levy AP, Jiang C, Kaelin WG, Jr., Goldberg MA. Negative regulation of hypoxia-inducible genes by the von Hippel-Lindau protein. Proceedings of the National Academy of Sciences of the United States of America. 1996;93(20):10595-9
  141. 141. Yang JC, Haworth L, Sherry RM, Hwu P, Schwartzentruber DJ, Topalian SL, et al. A randomized trial of bevacizumab, an anti-vascular endothelial growth factor antibody, for metastatic renal cancer. The New England journal of medicine. 2003;349(5):427-34
  142. 142. Choueiri TK, Motzer RJ. Systemic Therapy for Metastatic Renal-Cell Carcinoma. The New England journal of medicine. 2017;376(4):354-66
  143. 143. Krzywinska E, Stockmann C. Hypoxia, Metabolism and Immune Cell Function. Biomedicines. 2018;6(2):56
  144. 144. Messai Y, Gad S, Noman MZ, Le Teuff G, Couve S, Janji B, et al. Renal Cell Carcinoma Programmed Death-ligand 1, a New Direct Target of Hypoxia-inducible Factor-2 Alpha, is Regulated by von Hippel-Lindau Gene Mutation Status. Eur Urol. 2015;70(4):623-32
  145. 145. Pawlus MR, Wang L, Hu CJ. STAT3 and HIF1alpha cooperatively activate HIF1 target genes in MDA-MB-231 and RCC4 cells. Oncogene. 2014;33(13):1670-9
  146. 146. Battelli C, Cho DC. mTOR inhibitors in renal cell carcinoma. Therapy. 2011;8(4):359-67
  147. 147. Tian T, Li X, Zhang J. mTOR Signaling in Cancer and mTOR Inhibitors in Solid Tumor Targeting Therapy. International journal of molecular sciences. 2019;20(3):755
  148. 148. Wu D, Su X, Lu J, Li S, Hood BL, Vasile S, et al. Bidirectional modulation of HIF-2 activity through chemical ligands. Nature chemical biology. 2019;15(4):367-76
  149. 149. Cho H, Du X, Rizzi JP, Liberzon E, Chakraborty AA, Gao W, et al. On-target efficacy of a HIF-2alpha antagonist in preclinical kidney cancer models. Nature. 2016;539(7627):107-11
  150. 150. Chen W, Hill H, Christie A, Kim MS, Holloman E, Pavia-Jimenez A, et al. Targeting renal cell carcinoma with a HIF-2 antagonist. Nature. 2016;539(7627):112-7
  151. 151. Wallace EM, Rizzi JP, Han G, Wehn PM, Cao Z, Du X, et al. A Small-Molecule Antagonist of HIF2alpha Is Efficacious in Preclinical Models of Renal Cell Carcinoma. Cancer research. 2016;76(18):5491-500

Written By

Olivia Lombardi and David Robert Mole

Submitted: 01 February 2021 Reviewed: 09 February 2021 Published: 19 March 2021