Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Oct 2.
Published in final edited form as: Biochim Biophys Acta. 2012 May 10;1826(2):272–296. doi: 10.1016/j.bbcan.2012.04.008

The biological kinship of hypoxia with CSC and EMT and their relationship with deregulated expression of miRNAs and tumor aggressiveness

Bin Bao a, Asfar S Azmi a, Shadan Ali a, Aamir Ahmad a, Yiwei Li a, Sanjeev Banerjee a, Dejuan Kong a, Fazlul H Sarkar a,b,*
PMCID: PMC3788359  NIHMSID: NIHMS512312  PMID: 22579961

Abstract

Hypoxia is one of the fundamental biological phenomena that are intricately associated with the development and aggressiveness of a variety of solid tumors. Hypoxia-inducible factors (HIF) function as a master transcription factor, which regulates hypoxia responsive genes and has been recognized to play critical roles in tumor invasion, metastasis, and chemo-radiation resistance, and contributes to increased cell proliferation, survival, angiogenesis and metastasis. Therefore, tumor hypoxia with deregulated expression of HIF and its biological consequence lead to poor prognosis of patients diagnosed with solid tumors, resulting in higher mortality, suggesting that understanding of the molecular relationship of hypoxia with other cellular features of tumor aggressiveness would be invaluable for developing newer targeted therapy for solid tumors. It has been well recognized that cancer stem cells (CSCs) and epithelial-to-mesenchymal transition (EMT) phenotypic cells are associated with therapeutic resistance and contribute to aggressive tumor growth, invasion, metastasis and believed to be the cause of tumor recurrence. Interestingly, hypoxia and HIF signaling pathway are known to play an important role in the regulation and sustenance of CSCs and EMT phenotype. However, the molecular relationship between HIF signaling pathway with the biology of CSCs and EMT remains unclear although NF-κB, PI3K/Akt/mTOR, Notch, Wnt/β-catenin, and Hedgehog signaling pathways have been recognized as important regulators of CSCs and EMT. In this article, we will discuss the state of our knowledge on the role of HIF-hypoxia signaling pathway and its kinship with CSCs and EMT within the tumor microenvironment. We will also discuss the potential role of hypoxia-induced microRNAs (miRNAs) in tumor development and aggressiveness, and finally discuss the potential effects of nutraceuticals on the biology of CSCs and EMT in the context of tumor hypoxia.

Keywords: Hypoxia, HIF, CSC, EMT, miRNAs

1. Introduction

Low concentrations of oxygen in cells or tissues, referred to as hypoxia, are one of the most pervasive microenvironmental stresses and are recognized as the most common features of solid tumors. Hypoxia has been known to be associated with many aspects of biological processes during tumor development and progression such as cell survival, invasion, angiogenesis, and cellular metabolic alterations. Clinically, hypoxia and its signaling pathway have been shown to be associated with resistance to radiotherapy and chemotherapy, contributing to increased risk of tumor recurrence and metastasis, leading to reduced overall survival rate and increased mortality [1-5]. It is estimated that up to 60% of locally advanced solid tumors exhibit hypoxic (1% O2 or less, compared to 2–9% O2 in the adjacent tissues) and/or anoxic (<0.01% O2, or no detectable oxygen) conditions throughout the whole tumor tissues [6]. Transient hypoxia is related to inadequate blood supply while chronic or prolonged hypoxia is related to increased oxygen diffusion distance due to tumor expansion. Both types of hypoxic conditions are associated with poor outcome of patients diagnosed with solid tumors.

Tumor hypoxia is usually induced by several microenvironmental factors such as inadequate vascularization due to tumor angiogenesis leading to aberrant vessels with altered perfusion; an increase in oxygen diffusion distances due to rapid expansion of tumor cells; and tumor- or therapy-associated anemia leading to the reduced capacity of oxygen transportation [6,7]. Tumor cells usually have a greater capacity to adapt to a hostile, hypoxic environment for survival, compared to normal cells, which contributes to their malignant and aggressive behavior. This adaptation is controlled by many factors, including transcriptional and post-transcriptional changes in gene expression. It has been estimated that up to 1.5% of the human genome are responsive to hypoxia at transcriptional levels [6,7]. Hypoxia inducible factors (HIF) are one of the most critical transcriptional regulators that mediate the adaptation of tumor cells to a hypoxic tumor microenvironment.

In this review, we will discuss the role of HIF signaling pathways in the maintenance of hypoxic cellular response within the tumor microenvironment, and will discuss the role of hypoxia and its signaling pathway in the regulation and sustenance of the phenotypes of cancer stem cells (CSCs) and epithelial-to-mesenchymal transition (EMT). Moreover, we will summarize the molecular interrelation between HIF signaling pathway and other known cellular signaling pathways such as nuclear factor of κB (NF-κB), phosphatidylinositol 3-kinase (PI3K)/Akt (protein kinase B)/mammalian target of rapamycin (mTOR), Notch, Wnt/β-catenin, and Hedgehog in defining tumor aggressiveness. We will also discuss the state of our knowledge on the potential role of hypoxia-mediated microRNAs (miRNAs) in tumor development and aggressiveness. Finally, we will discuss the state of our knowledge on the role of natural product-derived agents (nutraceuticals) as potential molecular regulators of hypoxia-associated biology of tumor aggressiveness as a potential therapeutics.

2. Hypoxia, HIF, and clinical prognosis in tumor

HIF proteins are the master transcriptional regulators of the cellular response to hypoxic tumor microenvironment [8,9], which are involved in the regulation of many key aspects of tumor development and progression. For example, cell proliferation and survival are mediated through the regulation of the gene expression of HIF downstream targets, cyclin-dependent kinase inhibitor 1A (CDKN1A) and B-cell lymphoma 2 (Bcl2)/adenovirus E1B 19 kDa protein-interacting protein 3 (BNIP3), a cell death inducer during hypoxia; adaptive cellular metabolism mediated through the regulation of glucose transporter 1 (GLUT1), GLUT3, lactate dehydrogenase A (LDHA) and pyruvate dehydrogenase kinase 1 (PDK1); microenvironmental acidity mediated through the regulation of carbonic anhydrase 9 (CAIX); invasion and metastasis through the regulation of C–X–C chemokine receptor type 4 (CXCR4), mesenchymal–epithelial transition factor (c-MET), matrix metalloproteinases (MMP), and lysyl oxidase (LOX); angiogenesis through vascular endothelial growth factor (VEGF), VEGFR1, and angiopoietin-2 (Ang-2); and stem cell maintenance via the regulation of octamer-binding transcription factor 4 (Oct4) [6,10-13]. The function of these HIF downstream target genes are reviewed elsewhere [14,15], and thus these are not the focus of this article.

A large number of clinical evidence suggest that HIF and its downstream targets are considered as key markers of clinical prognosis of patients diagnosed with solid tumors. Increased expression of HIF-1α has been identified to be associated with poorer prognosis with decreased disease-free survival in several early studies, which has been confirmed by a recent meta-analysis report [3]. Increased expression of VEGF and/or HIF-1α has been shown to be associated with poor prognosis [6,16-19]. The up-regulation of CAIX has also been associated with aggressive features with poor overall and relapse-free survival, consistent with the expression of HIF-1α [20-26]. Both markers are shown to correlate with both primary breast tumor and lymph node metastasis [26,27]. The up-regulation of GLUT1 and lactate dehydrogenase 5 (LDH-5) has been shown to be associated with poor prognosis, consistent with the expression of HIF-1α in many solid tumors [19,26-34]. High expression of BNIP3 in tumors is also reported to be associated with poor prognosis with increased risk of recurrence and decreased disease-free survival [27,35,36], and may be considered as independent prognostic factor for overall survival [27,35-38].

Recently, the expression of HIF-2α or concomitant with the expression of HIF-1α and its downstream targets, VEGF, Oct4, and erythropoietin (EPO), has been shown to be positively associated with poorer prognosis, increased rate of local recurrence, and reduced overall survival rate in various cancers [39-42]. These data clearly suggest that hypoxia and HIF signaling pathway play important roles in tumor development and aggressiveness.

3. Hypoxia, HIF, and treatment resistance in tumor

Hypoxia and HIF pathway have been considered as a negative factor for tumor therapy, and have been identified to be associated with the resistance to radiotherapy and chemotherapy [4,5]. Several clinical studies demonstrated that HIF-1α, and its downstream targets, CAIX and VEGF have been associated with resistance to chemotherapy [5,23,43-45], consistent with multiple recent findings [46-50], indicating that hypoxia is associated with chemotherapy resistance.

The relationship between hypoxia and resistance to radiation therapy has also been documented. A recent meta-analysis report in head and neck cancers suggests that hypoxic modification improves tumor control and survival in conjunction with curative intended radiation therapy of head and neck cancers [51]. Another recent meta-analysis report demonstrates that biological markers involved in angiogenesis and hypoxia are associated with poor prognosis of cervical cancer with chemotherapy and radiation therapy [52]. These clinical data suggest that targeting these hypoxia-induced signaling pathways in combination with chemo-radiation therapy may improve survival of patients diagnosed with advanced-stage cervical cancer.

Several experimental studies have confirmed the important role of hypoxia and HIF signaling pathway in the induction of treatment resistance. High expression of HIF-1 enhances the resistance to radiation as assessed by clonogenic survival assay under normoxic and hypoxic conditions [53], whereas inactivation of HIF-1 signaling pathway by HIF-1α inhibitors or siRNA has been shown to increase the sensitivity of tumor to radiation therapy in vitro and in vivo [13,53,54]. Other experimental studies also revealed that HIF-1α deficiency in various cancer cells, including CSCs, are more susceptible to chemotherapeutics and ionizing radiation therapy than its wild type cells [15,45,55]. These findings are consistent with other findings [56-59], which clearly support that hypoxia-induced signaling pathway plays an important role in both de novo (intrinsic) and acquired (extrinsic) resistance to conventional therapeutics in most solid tumors.

Although the exact molecular mechanism(s) of how hypoxia and HIF pathway lead to therapeutic resistance is not fully understood, several mechanisms have been proposed, such as decreased cytotoxicity due to the lack of oxidation of DNA free radicals by hypoxia, hypoxia-induced cell cycle arrest, compromised drug delivery due to increased distances of tumor cells to aberrant vasculature, impaired DNA repair system, hypoxia-induced extracellular acidification, and altered cellular metabolisms through multiple signaling pathways such as GLUT1 and ATP-binding cassette transporter G2 (ABCG2), a known drug resistant marker, which affects drug efficacy. It has also been considered that hypoxia-induced development and maintenance of CSC and EMT phenotypes would contribute to increased therapeutic resistance. Recently, hypoxia-induced VEGF has been shown to increase treatment resistance, due to the protective role of VEGF against cytotoxic effects of radiation in endothelial cells [60], further suggesting mechanistic role of VEGF in resistance to chemotherapy and radiation therapy [6,61].

4. Hypoxia-inducible signaling pathways in tumor

Emerging evidence suggests that hypoxic-inducible signaling pathways are significantly associated with the development and progression of tumors. As the major transcription regulators responsible for hypoxic stress, HIFs have been demonstrated to play a pivotal role in tumor growth, angiogenesis, invasion, and metastasis (Fig. 1). To date, three HIF family proteins, namely HIF 1, 2, and 3 have been identified in humans. We will discuss the role of these HIF proteins in tumorigenesis as documented in the following paragraphs.

Fig. 1.

Fig. 1

Hypoxia driven tumor angiogenesis. Hypoxic tumor microenvironment enhanced through desmoplasmic stroma results in activation of HIF proteins that influences aberrant miRNA expression dependent angiogenic signaling resulting in tumor angiogenesis.

4.1. The role of HIF-1α signaling in tumor

HIF, a member of per-aryl hydrocarbon receptor nuclear translocator (ARNT)-sim (PAS) family of heterodimeric basic helix–loop–helix (bHLH) transcription factor, are composed of α subunit and β subunit [62-67]. Both α and β subunits belong to bHLH/PAS domain transcriptional factors [66]. To date, there are three subunits of α, namely, 1α, 2α, and 3α that have been demonstrated. However, HIF-1α/β is reportedly the most common heterodimer involved in hypoxic response [10,68-70]. The HIF-1β subunit (alsoknownasARNTorARNT2) is identical to the arylhydrocarbon receptor nuclear translocator, which serves as a heterodimeric partner with the arylhydrocarbon receptor. HIF-1β is constitutively expressed and present in the nucleus across many cell types, and is regulated in oxygen-independent pattern [71-73]. The in vivo experimental studies have revealed that functional loss of HIF-1β suppresses tumor growth, consistent with decreased expression of VEGF and GLUT3 in hepatoma xenograft models [74,75], suggesting its potential role in tumorigenesis.

Expression of α subunit is regulated in an oxygen-dependent manner and it is induced by cellular hypoxia. Its expression is usually maintained at low levels in most cells with normal oxygen tension [67]. The levels of HIF-1α are the primary determinant of HIF-1 DNA binding and transcriptional activity [76]. The induction of HIF- 1α is a critical step in the hypoxic response which occurs through the up-regulation of its mRNA expression, protein stabilization, and nuclear localization. Under normoxic condition, post-translational hydroxylation of proline residues at the positions of 402 and 564 (Pro402 and Pro564) in the oxygen-dependent degradation domain (ODD) of HIF-1α is induced by oxygen sensor dioxygenase proteins and prolyl hydroxylases (PHD 1, 2, 3), especially PHD2 [77-80]. The hydroxylated HIF-1α binds to tumor suppressor protein the von Hippel–Lindau (VHL), an E3 recognition component of the ubiquitin ligase complex by these hydroxylated groups of HIF-1α [81-84]. These interactions lead to rapid degradation of HIF-1α through a ubiquitin and proteasome-dependent pathway [83,85-90], leading to the inactivation HIF-1α [88,91]. This hydroxylation of HIF-1α occurs quite rapidly under normoxic conditions, resulting in very rapid HIF-1α turnover, resulting in maintaining the minimal levels of HIF 1α activity [92-94]; however, this regulation is altered under hypoxic condition with the tumor microenvironment as discussed later. In addition to proline hydroxylation, asparagine at the position of 803 (Asn803) in the COOH-terminal transactivation domain (CAD) of HIF-1α can be hydroxylated by the factor-inhibiting HIF (FIH), an oxygen sensor hydroxylase enzyme under normoxic condition [77-80,95]. These modifications inhibit HIF transcriptional activity by preventing the interaction between CAD of HIF-1α and p300/CBP transcriptional co-activator. FIH also can block the interaction between HIF-1α and its co-activators, leading to the inactivation of HIF-1α transcriptional activity. Under hypoxic condition, HIF-1α is stabilized due to hypoxia-induced inhibition of PHD and FIH activities, and then it is translocated into the nucleus where it dimerizes with the constitutively expressed HIF-1β [77,92]. The HIF-1α/β heterodimers bind to its cognate transcriptional enhancer element, referred to as hypoxia-response element (HRE) (5′-RCGTG-3′, where R is A or G), in the promoter/enhancer regions of a number of HIF target genes in the presence of p300/CREB binding protein (CBP) coactivators [96-98], leading to their transcriptional activation. These HIF responsive gene targets include glucose transporters, glycolytic enzymes, angiogenic growth factors, and several molecules involved in apoptosis and cell proliferation such as GLUT1,3, EPO, transferrin, endothelin-1 (ET-1), inducible nitric synthase (iNOS), heme oxygenase 1, PDK1, lactate dehydrogenase A (LDHA), BNIP3, VEGF, insulin-like growth factor (IGF) and IGF-binding proteins [99,100]. Currently, more than 60 HIF target genes have been identified in human cells, as reviewed elsewhere [13,14,101,102].

VHL plays an important role in the regulation of HIF signaling pathway, and is negatively associated with HIF pathway. Functional loss or mutation of VHL increases the activation of HIF signal pathway. Clinically, mutation in this gene causes VHL disease, congenital polycythemia, and several sporadic tumor types. Several other tumor suppressors such as p53, phosphatase and tensin homolog deleted on chromosome 10 (PTEN), endogenous Akt inhibitor protein, tuberous sclerosis 2 (TSC2) and liver kinase B1 (LKB1) have been shown to regulate the activation of HIF-1α. Functional loss or mutation of these genes induces activation of HIF-1 activation even under normoxic conditions [13,103-106].

A number of experimental studies have demonstrated that increased activation of HIF-1α has a role in promoting cell survival, invasion, angiogenesis, and tumor growth [107]. HIF-1α is expressed in a wide range of human cancer cells. High expression of HIF-1α has been identified in many human cancers including brain, breast, colon, lung, pancreatic, prostate and ovarian cancer [10,11,55,100,108,109], and is clinically associated with poorer prognosis of many tumors.

4.2. The role of HIF-2α signaling in tumor

HIF-2α (also known as endothelial PAS domain protein 1, EPAS1) has been shown to share transcriptional target genes with HIF-1α such as VEGF, tyrosine kinase receptor with immunoglobulin and epidermal growth factor homology 2 (Tie2), Ang-2 and fms-like tyrosine kinase 1 (Flt-1), and also regulated by cellular oxygen tension. HIF-1α and HIF-2α have the homologous DNA-binding sequence binding sites in the promoter/enhancer regions of their target genes [110]. Similar to HIF-1α, HIF-2α is stabilized at a wider range of oxygen tensions, ranging from severe hypoxia (<1% oxygen) to more physiologically relevant tension in tumors (2–5% oxygen) [111], and activates downstream transcriptional targets. It has been documented that HIF-2α has a similar role in the induction of angiogenesis via upregulation of VEGF production, and may play a dominant role in tumor angiogenesis [112]. Despite their similarities, HIF-2α expression has been shown to be restricted to endothelial cells of vascular organs and had several unique transcriptional targets such as Oct4 and transforming growth factor-α (TGF-α). These targets are different from the canonical pro-angiogenic hypoxic response, which suggests an important and specific role for HIF-2α in the regulation of other cellular processes such as pluripotency.

The role of HIF-2α in HIF signaling pathway is not fully understood. High expression of HIF-2α has been recently identified in certain cancer cells including CSCs. Emerging evidence from experimental studies suggests that HIF-2α plays an important role in tumorigenesis, for example, inhibition of HIF-2α by its siRNA suppresses the growth of tumor derived from pVHL-defective renal carcinoma cells in vivo [113]. One recent experimental study paradoxically showed that HIF-2α over-expression can contribute to the progression of non-small cell lung cancer (NSCLC), whereas HIF-2α deletion promotes Kras-driven lung tumor development [114]. However, multiple recent experimental studies have revealed that HIF-2α is required for maintaining the phenotype and function of various stem cells such as glioblastoma, neuroblastoma, renal cancer, and non-small cell lung cancer [111,115-120]. The clinical data shows that there is a concomitant increased expression of HIF-1α and HIF-2α expression in certain cancers [112,121-123], and that the expression of HIF-2α and/or HIF-1α as well as its downstream targets has been shown to be associated with poorer prognosis of malignant diseases [115,122,124-127]. It appears that there is a crosstalk between HIF- 1α and HIF-2α in tumor development and progression. Further molecular and mechanistic investigations are needed in order to fully elucidate the role of HIF proteins in human solid tumors.

4.3. The role of HIF-3α signaling in tumor

The function and regulation of HIF-3α remain unclear. Limited studies have shown that HIF-3α has several splice variants, and its limited expression pattern has been identified in the eye and in the cerebellum, kidney, heart and lung [128-131]. Some HIF-3α variants have been shown to be a dominant-negative regulator of the HIF-1α and 2α isoforms, and are thought to be direct transcriptional targets of HIF-1α activity under hypoxic conditions [130,131]. It has also been shown in vivo that HIF-3α can be induced by insulin-induced hypoglycemia under normoxic conditions in cerebral cortex and kidney [129]. One recent study reveals that the expression of HIF-3α isoforms is cell type specific in vascular cells; however, only one isoform, HIF-3α2, is mediated by hypoxia and regulated by NF-κB. Overexpression of HIF-3α isoforms decreases hypoxia-induced expression of VEGF and enolase 2, which confirms the notion of a hypoxia-inhibition of HIF-3α as a negative regulator of HIF signaling pathway [132].

The role of HIF-3α and its variants in tumorigenesis has not been optimally exploited. One experimental study has shown that HIF-3α4, an alternatively spliced variant of human HIF-3α with similar domain structure as the murine inhibitory PAS protein (IPAS), is able to form an abortive transcriptional complex with HIF-2α and prevents the engagement of HIF-2 to the HRE site located in the promoter/enhancer regions of hypoxia-inducible genes. In addition, the re-expression of HIF-3α4 in VHL-null 786-O renal cancer cells via adenovirus decreases the endogenous expression of HIF-2 downstream target gene expression and suppresses the growth of 786-O tumor xenografts in severe combined immunodeficiency (SCID) mice. These data suggest that HIF-3α isoforms may have the tumor suppressive function [133].

5. The regulatory role of hypoxia and HIF pathway in CSCs

A number of studies have provided supportive evidence that hypoxia and HIF signaling pathway play pivotal roles in the regulation of the phenotype and function of CSCs, contributing to tumor aggressiveness. The CSCs constitute a sub-population of malignant cells within a tumor mass and posses the ability to self-renew giving rise to heterogeneous tumor cell populations with a complex set of differentiated tumor cells [134-137]. The theory of CSCs has fundamental clinical implications especially because CSCs have been identified in many malignant tumor tissues and the CSCs are considered to be highly resistant to chemo-radiation therapy [134-138] relative to differentiated daughter cells. This provides an explanation for the clinical observation that tumor regression alone may not correlate with patient survival [138] because of tumor recurrence due to the presence of CSCs. The CSCs have been believed to play critical roles in treatment resistance and cancer metastasis and recurrence after conventional therapy through multiple mechanisms and networks, as reviewed by us recently [137,139,140].

Hypoxia has been recognized as an important factor that regulates the sub-population of normal stem cells and maintains the normal tissues or non-stem cell tissues in a stem cell state during embryonic and adult development [98,141,142]. Emerging experimental evidence indicates that hypoxia also plays an important role in the regulation of the phenotype and function of CSCs by enhancing the self-renewal capacity and maintenance of undifferentiated state of CSCs [143-145], consistent with recent findings in glioma CSCs [146,147]. Hypoxic CSCs robustly retain the undifferentiated phenotype, whereas normoxic CSCs retained differentiated state. The hypoxia-induced phenotype of CSCs is reversible when it re-grows at normoxic conditions [148]. It has been shown that hypoxia is able to maintain the stem-like phenotypes in neuroblastomas and activate signaling pathways that are associated with undifferentiated phenotypes of normal stem cells, including sex determining region Y box 2 (Sox2), Oct4 and Notch-1 signaling [118]. The hypoxia-induced activation of the signature genes of CSCs may be one of the major reasons that hypoxia is associated with increased tumor aggressiveness leading to poorer clinical outcome (Fig. 2).

Fig. 2.

Fig. 2

Hypoxia as a driver of aggressive tumor phenotype. HIF (HIF-1α and HIF-2α) activation drives two critical tumor promoting pathways (1) increase in CSC markers (Oct4, Nanog, Sox2 and Snail) and (2) EMT promoting markers (ZEB2, Snail, Twist, Wnt, Slug, Notch and TGF-β). These HIF driven pathways interact with one another leading to enrichment/enhancement in both CSC and EMT cell populations that in turn is a driver of aggressive tumor type.

Emerging evidence suggests that hypoxic areas within a tumor or the areas of necrotic tumor tissues are considered as a niche where CSCs reside [149]. Such hypoxic niche for CSCs may have an important role in the development and progression of tumors. Therefore, it is possible that tumors may evolve and develop from mutation in normal stem cells or from non-stem cell population during hypoxic conditions. The precise mechanism by which hypoxia mediates the phenotype and function of CSCs is not fully understood. It has been documented that hypoxia-mediated phenotype and function of CSCs is regulated by HIF proteins, specifically HIF-1α- and HIF-2α-mediated signaling pathways, which participates in the transcriptional activation and regulation of HIF targets, such as Oct4 and other CSC markers, as reviewed elsewhere [148,150].

As stated earlier that the expression of both HIF-1α and HIF-2α are associated with tumor aggressiveness. A number of experimental studies have clearly provided convincing evidence showing that both HIF-1α and HIF-2α plays important roles in the regulation of CSC phenotypes and functions, in a cell type specific manner. For examples, several recent studies have demonstrated that hypoxia-induced HIF-1α promotes the phenotype and function of CSCs, consistent with the up-regulation of HIF-1α, and HIF-target genes such as Oct4, Nanog, c-Myc, Notch-1, and CD133, the CSC markers, in various cancer cells including CSCs [144,146,151-154]. Conditional knock-down of HIF-1α attenuates the expansion of hypoxia-induced CD133+ CSCs, which are confirmed by other recent findings [155,156]. These results collectively suggest that HIF-1α is required for the maintenance of the phenotype and function of CSCs in certain cancers.

Other recent investigations were focused on the role of HIF-2α in the regulation of CSCs. It has been shown that HIF-2α is highly expressed in CSCs; also known as tumor initiating cells (TIC) of neuroblastoma, glioblastoma, renal cell carcinoma, and breast carcinoma are associated with unfavorable disease outcome [119,120,157]. The expression of HIF-2α can be induced by hypoxia, and increased expression of CD133 in CSCs [158] and the tumorigenic capacity of glioma stem cells [117]. Conditional loss of HIF-2α increases vessel abnormalities, decreases significantly the proliferation and selfrenewal capacity of CSCs, and suppresses tumor growth and metastasis [159,160]. Forced over-expression of HIF-2α induces stem cell markers Oct4, Nanog, Sox2, and c-Myc, and augments the tumorigenic capacity of the non-stem cell population [159]. It is interesting to note that HIF-2α may participate in the up-regulation of acidity-induced CSC phenotype and function in an oxygen-independent manner [161]. These findings strongly suggest that HIF-2α is required for the phenotype and function of CSCs.

The role of differential expression of HIF-1α and HIF-2α between non-stem cells and CSCs has paid significant attention in recent years, which may depend on the degree and period of hypoxia as well as a specific cell type dependent phenomenon. It has also been shown that HIF-2α is only significantly present in the population of CSCs, and remarkably induced by hypoxia [117,162]. In comparison, HIF-1α is present in both stem and non-stem tumor cells and is only stabilized under more severe hypoxic conditions [117], suggesting the differential roles of these two isoforms in CSCs.

It has been postulated that HIF-1α involves in the adaptation to acute/transient, and HIF-2α to chronic/prolonged hypoxia [115,163]. A recent experimental data has shown that hypoxia associated factor (HAF, also known as SART1), a potent E3 ubiquitin ligase, is highly expressed in various cancer cells and its levels are decreased during acute hypoxia, but increased following prolonged hypoxia [164,165]. HAF can inhibit HIF-1α activity through its direct binding to HIF-1α protein for ubiquitination and degradation in an oxygen- and pVHL-independent mechanism. It is also known that HAF binds to HIF-2α, but at a different site than HIF-1α, and increases HIF-2α transactivation without causing its degradation. Thus, HAF switches the hypoxic response of the cancer cell from HIF-1α-dependent to HIF-2α-dependent transcription and activates HIF target genes such as MMP-9, plasminogen activator inhibitor 1 (PAI-1), and Oct-3/4. The switch to HIF-2α-dependent gene expression caused by HAF also promotes the expansion of tumor stem cell population, resulting in highly aggressive tumors in vivo [164,165]. These results suggest the differential expression and roles of HIF-1α and HIF-2α in the regulation of CSCs, which could indeed be dependent on hypoxic tumor microenvironment. However, the exact molecular mechanism(s) of how HIH-1α and HIF-2α interplay their roles in the regulation of CSC require further in-depth investigation.

6. The regulatory role of hypoxia and HIF pathway in EMT

The acquisition of EMT during tumor development progression is known to play important roles, which is also associated with the formation of the primary mesoderm from upper epiblast epithelium during neural crest cell formation from part of the ectoderm, and in palatal formation during embryonic development [139,166]. The EMT also occurs during adult placenta formation, and in the formation of fibroblasts during inflammation and wound healing after birth [139,166]. EMT is an important biological process by which epithelial-like cells with a cobblestone phenotype acquire mesenchymal characteristics with a spindle-shaped fibroblast-like morphology in human tumors. This process involves a disassembly of cell–cell junctions, such as down-regulation and relocation of E-cadherin and zonula occludens-1 (ZO-1), which are epithelial cell phenotype markers, and down-regulation and translocation of β-catenin from the cellular membrane to the nucleus, re-organization of actin cytoskeleton, and up-regulation of mesenchymal cell phenotype markers (such as vimentin, fibronectin, and N-cadherin). This process confers mesenchymal phenotypic cells to have less cell adhesion capacity, which leads to increased cell migration and invasion, resulting in tumor aggressiveness [167-170].

During the acquisition of EMT phenotype, several transcription factors such as zinc-finger E-box binding homeobox 1 (ZEB1) and ZEB2 (also known as SIP2, septin interacting protein 2), and Snail1 (also known as Snail), Snail2 (also known as Slug), Twist1 (also known as Twist), and E47 (also known as T cell factor, TCF3) have been shown to be critical mediators of EMT phenotype induced by a variety of inducers in different cell lines including cancer cells [170,171]. It has been demonstrated that ZEB1 regulates the expression of genes by binding to ZEB-type E-boxes (CACCTG) within the promoter region of target genes, resulting in chromatin condensation and gene inactivation [170,172]. The expression of E-cadherin, a marker for epithelial cell phenotype is repressed by ZEB1 through its binding to ZEB-type E-boxes in the E-cadherin gene promoter, which is fundamental for the acquisition of EMT phenotype [173]. Although EMT has been originally described in embryonic development, where cell migration and tissue remodeling have a primary role in the regulation of morphogenesis in multi-cellular organisms, recent experimental data and clinical findings have provided supportive evidence showing that EMT plays a pivotal role in tumor aggressiveness mediated via induction of cancer cell invasion and migration, and the phenotype and function of CSCs, thereby leading resistance to chemo- and radiation therapy, resulting in tumor recurrence [174].

A large number of data demonstrate that hypoxia induces EMT characteristics in a variety of cells including cancer cells [169,175,176]. Moreover, the hypoxia-induced EMT is tightly mediated by HIF-signaling pathway, which contributes to aggressive tumor growth and invasiveness. For example, several experimental studies have demonstrated that growth under hypoxic condition reprogram epithelial cells to a more mesenchymal phenotype, due to the activation of E-cadherin transcription repressors, leading to the promotion of tumor invasive potential [177-180], consistent with more recent findings [176]. A complex molecular crosstalk between hypoxia-induced pathway and EMT has not yet fully understood. Several aspects of potential molecular mechanisms have been proposed, as reviewed recently [170]. First, the activation of HIF-1 and 2 can induce EMT by up-regulation of EMT-associated transcription factors or repressors such as Twist, Snail, Slug, and SIP1/ZEB2 in many different cancer cells [179,181-183], as supported by more recent findings [184]. Secondly, hypoxia and HIF pathway activate the EMT-associated signaling pathways such as TGF-β, Notch, NF-κB, Wnt/β-catenin, and Hedgehog [173,178,185-213], Next, hypoxia and HIF pathway can induce EMT phenotype or characteristics by the regulation of EMT-associated inflammatory cytokines such as increased expression of hypoxia-induced tumor necrosis factor α (TNF-α), interleukin 6 (IL-6), and IL-1β, which promote the induction of EMT phenotype [214-216]. Finally, hypoxia and HIF pathway can induce EMT by direct or indirect regulation of proteins or enzymes that mediate cell–matrix interactions and facilitate motility and invasion mediated through the regulation of LOX/LOX2, Hey1, Hes1, and urokinase-type plasminogen activator (uPA) expression [217-219]. The data from recent studies suggest that other cellular signaling pathways such as VEGF and epigenetic regulators have been shown to play important roles in the hypoxia-induce EMT in cancer cells [220-223]. For example, hypoxia-induced or HIF-induced histone deacelylase 3 (HDCA3) is reported to be required for hypoxia-induced EMT phenotype [224]. One in vitro and in vivo study revealed that HIF-2α promotes the expression of several genes associated with EMT phenotype markers such as ZEB1, Snail, SIP1, and vimentin, and increases non-small cell lung tumor growth in xenograft mouse model [116]. Therefore, targeting both HIF pathway and EMT could provide a therapeutic strategy in the future (Fig. 2).

7. The role of VEGF signaling in tumor

Angiogenesis is classically defined as the formation of new blood vessels from the existing vascular bed, which is an essential biological process during development and diseases. Angiogenesis has been considered as one of the hallmarks of cancer, and plays an essential role in tumor growth, invasion and metastasis. Tumor angiogenesis is induced by hypoxic condition through the regulation of several pro-angiogenic factors including VEGF, one of the major HIF target gene products. The VEGF is a well-known potent endothelial specific mitogen and has a central role in tumor angiogenesis [13,70,102,225]. VEGF and its receptor signaling pathway is one of the major signaling pathways and is best characterized. The increased expression of VEGF has been identified to be associated with a wide range of solid tumors, as stated earlier. Moreover, VEGF has been considered as negative predictor of tumor-related outcome.

In addition to a central role in tumor angiogenesis, emerging evidence suggests that VEGF might have other regulatory roles in tumor aggressiveness by angiogenesis-independent mechanism. Several studies have demonstrated that VEGF or VEGFR-1 is able to inhibit epithelial cell phenotype, and to mediate EMT phenotype in various cells including pancreatic cancer cells through the regulation of EMT-associated transcription factors, Snail, Twist, and Slug [221-223]. Additional studies have confirmed that the activation of VEGF or VEGFR may induce EMT phenotype in cancer cells [220,226,227]. These results suggest that hypoxia-induced VEGF signaling pathway not only acts as a regulator of angiogenesis, but also acts as a potential inducer of EMT, both of which contributes to tumor aggressiveness.

The role of VEGF in the regulation of the phenotype and function of CSCs has not been fully investigated. A recent report has shown that VEGF promotes cardiovascular stem cell migration via PI3K/Akt pathway in a concentration-dependent manner under hypoxic conditions [228]. The VEGF receptor inhibitor SU5416 or the PI3K/Akt inhibitor wortmannin suppresses the VEGF-induced cardiac stem cell accumulation and migration [228]. Another study have demonstrated that stem cell-like glioma cells isolated from human glioblastoma samples produce high levels of VEGF, significantly form highly-vascular and hemorrhagic tumor in the brains of immunocompromised mice. Treatment of these CSC-like cells with VEGF inhibitor bevacizumab inhibits their ability to induce endothelial cell migration and tube formation in vitro, and suppresses CSC-derived tumor growth [229], which suggests that hypoxia-induced VEGF expression may contribute to tumor initiating capacity through stem cells.

8. The role of NF-κB in HIF signaling pathway in tumor

HIF signaling pathway is essential for cellular response to hypoxia. However, almost 20 different transcription factors have displayed directly or indirectly hypoxic sensitivity, including NF-κB. The NF-κB has been identified to be activated by hypoxia for more than a decade in a wide variety of cells [230,231]. NF-κB is a family of hemo- or heterodimer transcription regulators including RelA (p65), RelB, c-Rel, NF-κB1 (p105/p50), and NF-κB2 (p100/p52), which are typically activated following stimulation of cells with pro-inflammatory cytokines such as TNF-α and IL-1β, antigens and bacterial liposaccharides (LPS), growth factors, UV light, and reactive oxygen species (ROS) [232,233]. Normally, NF-κB proteins are sequestered in the cytoplasm of non-stimulated cells by a family of inhibitory proteins of NF-κB (IκB-α and IκB-β), which bind to NF-κB and mask its nuclear translocation. Once cells are stimulated, IκB proteins become phosphorylated by the IκB kinase complex (IKK-α, β, and γ), and ubiquitinated, and degraded by the proteasomes, eventually leading to nuclear translocation and activation of NF-κB target gene transcription.

The NF-κB is a well-known critical transcription factor in a wide range of cancers that regulates the expression of NF-κB target genes associated with a variety of cellular processes such as cell survival, proliferation, invasion, migration, angiogenesis and tumor metastasis, leading to the resistance to chemo- and radio-therapy, tumor recurrence or relapse, and increased risk of poor prognosis [234-236]. Moreover, emerging evidence suggests that NF-κB also plays an important role in EMT and CSCs, contributing to tumor aggressiveness [237-240].

A large number of studies have indicated that hypoxia could activate NF-κB signaling pathway in a variety of cells including cancer cells. NF-κB is primarily a regulator of inflammatory and anti-apoptotic gene expression. It is reasonable to believe that NF-κB activation is required to inhibit apoptosis to enable a hypoxic cell to survive through the period of hypoxic insult. Hypoxic action of NF-κB signaling pathway also involves promoting hypoxic inflammatory response through the regulation of gene expression of pro-inflammatory cytokines such as TNF-α, IL-6, and IL-1β, as well as adhesion molecules, enzymes, and pro-inflammatory enzymes [241,242]. It is reasonable to speculate that hypoxia induced NF-κB activation may be contributing to the maintenance of CSCs and EMT during the development and progression of tumors especially because both HIF and NF-κB signaling pathways are known to enhance the induction of EMT phenotype and maintenance of stem cell phenotype and function in tumor microenvironment.

HIF-1α has been shown to be required for the activation of NF-κB [243]. The functional loss or deficiency of HIF-1α decreases NF-κB activation and p65 expression whereas increased expression of HIF-1α results in the activation of NF-κB through hyperphosphorylation of IκB and phosphorylation of p65 at Ser276 in keratinocytes [244]. The HIF-1α-mediated activation of NF-κB has been shown to be consistent with cancer cell survival, invasion, and tumor growth [240,243,245-247]. Therefore, hypoxic activation of HIF pathway plays an important role in the regulation of NF-κB pathway in tumor microenvironment. The molecular mechanism of how HIF pathway regulates NF-κB pathway has been partly exploited in several studies, which suggest that HIF pathway activates NF-κB pathway mainly through the regulation of IKK signaling-mediated regulation of canonical NF-κB signaling.

Some earlier studies has shown that hypoxia induces NF-κB activation through the regulation of the activation of IKK signaling, especially IKK-β [248], which was further confirmed by recent findings [59,247, 249-252]. Hypoxia-induced IKK activation has been considered to be mediated through PHD domain because IKK-β contains sequence-analogous motifs (LxxLAP motifs), which are known targets for hydroxylation by PHD1 [248]. Pharmacological or genetic inhibition of hydroxylases PHD1 signaling leads to increasedNF-κB signaling[248]. Recently, PHD3 is found to inhibit the phosphorylation of IKK-β and NF-κB proteins independent of its hydroxylase activity by blocking the interaction between IKK-β and heat shock protein 90 (Hsp90) that is required for phosphorylation of IKK-β in colorectal cancer cells. Knock-down of PHD3 increased IKK signaling and the resistance of these cancer cells to the effects of TNF-α and increased tumorigenesis [253]. Furthermore, two additional new studies have demonstrated that hypoxia-induced IKK kinases such as X-linked inhibitor of apoptosis protein (XIAP) and TGF-β activated kinase 1 (TAK1) could mediate IKK activity via IKK-β, leading to the activation of NF-κB [245,249]. Conditional deletion of XIAP, a potential E3 ubiquitin ligase, inhibits hypoxia-induced IKK activation. Over-expression of XIAP leads to increased phosphorylation of IKK in both normoxic and hypoxic conditions [249]. These findings suggest that hypoxia causes activation of IKK, specifically via IKK-β, and thus may have a key role in the regulation of NF-κB pathway under tumor hypoxic microenvironment.

Moreover, FIH is also considered to regulate hypoxic induction of NF-κB activation through the hydroxylation of NF-κB components such as p105 and IκB-α because these components of NF-κB contains ankyrin repeat domain, which are subject to hydroxylation by FIH [254]. However, the functional consequence of the hydroxylation of IκB-α by FIH is not conclusive [255,256]. Several studies have shown that hypoxic activation of NF-κB pathway is, at least, in part HIF-1-dependent [243,244,247]. It is speculated that HIF-1 may directly interact with NF-κB proteins, promoting DNA binding at the promoter region of NF-κB target genes.

In addition to HIF-mediated activation of NF-κB, it is known that NF-κB could regulate HIF pathway in tumorigenesis. A large number of experimental studies have clearly indicated that NF-κB could directly regulate HIF signaling pathway within the tumor microenvironment because HIF promoter contains active NF-κB binding site in position −178/−188 [257-259]. The mutation of NF-κB DNA binding site in the HIF-1α promoter region leads to the loss of hypoxia-induced HIF-1α upregulation [257,259]. Moreover, blocking individual NF-κB members by its specific siRNAs decreases HIF-1α activity [257,259]. The activation of NF-κB induced by TNF-α or p50, p65 results in increased levels of HIF-α mRNA and protein [257,259]. It has also been shown that knockdown of IKK-β decreases HIF-1α activity [252], consistent with other recent findings indicating that NF-κB up-regulates HIF-1 signaling pathway via IKK-β [260-262]. These results suggest that IKK signaling is required for the activation of HIF-1α activity. Non-hypoxic stimulation such as cytokines especially TNF-α and IL-4, also activate HIF-1α activity through NF-κB dependent mechanism [254,260,263]. Therefore, NF-κB regulates HIF-1 signaling pathway to maintain the basal levels of HIF-1α under normoxia condition, and the levels of HIF-1α increases under hypoxia.

Interestingly, one earlier study has shown that NF-κB induces HIF-2α activation through the interaction with IKK-γ (also known as NEMO, NF-κB essential modulator), leading to CBP/p300 recruitment for transcriptional activation of HIF-2α [264]. Moreover, one recent study has shown that NF-κB directly regulates HIF-1β mRNA and protein in Drosophila. NF-κB-mediated induction of HIF-1β results in the modulation of HIF-2α. The over-expression of HIF-1β can rescue HIF-2α protein levels following NF-κB depletion [265]. Furthermore, NF-κB has also been found to regulate HIF-3α, a negative regulator of HIF signaling pathway [132]. However, the role of NF-κB in the regulation of HIF-3α in the tumor microenvironment is not fully understood.

One recent clinical study revealed that the expression of HIF-1α is positively associated with the expression of RelA in gastric cancer patients. The in vitro experiment showed that over-expression of IκB-αM (supper-suppressive mutant form) suppresses tumor growth, angiogenesis, and HIF-1α expression [246], consistent with previous findings indicating that the interaction of HIF and NF-κB pathways contributes to breast cancer metastasis through the induction of EMT together with migration through p65-lysine acetylation and HDAC-dependent epigenetic mechanism to up-regulate NF-κB and HIF [266]. Taken together, the interaction between HIF and NF-κB signaling pathways plays a pivotal role in tumor aggressiveness. However, the precise molecular crosstalk between these pathways in the tumor microenvironment is complex and still requires in-depth investigation. In summary, NF-κB and HIF pathways appear to have an important role in the tumor microenvironment; however, the precise molecular mechanism of crosstalk between NF-κB and HIF require further in-depth investigation.

9. The role of PI3K/Akt/mTOR in HIF signaling pathway in tumor

PI3K/Akt/mTOR pathway has been known to play a key role in a wide variety of cellular processes including cell proliferation, adhesion, migration, invasion, metabolism, and survival [267]. Phosphorylation of PI3K by microenvironment stimuli such as growth factors, cytokines, and hormones results in the activation of Akt, which leads to the phosphorylation of mTOR (an active form). The activation of mTOR complex (mTORC1 and mTORC2) then phosphorylates p70- S6 kinase (S6K1) and eukaryotic initiation factor 4E (elF4E) binding protein 1 (4-BP1), via mTORC2, finally leading to the up-regulation of gene transcription and translation [268-270].

The tumor suppressor protein, PTEN, a phosphatase opposes the activation of PI3K, thereby down-regulates the level of activated Akt, leading to the reduction of mTOR signaling [268,271,272]. Aberrant activation of PI3K/Akt/mTOR signaling pathway has been identified in many tumors and has been considered to be associated with tumor aggressiveness, in part, mediated through the regulation of multiple cellular signaling pathways involving in the induction of EMT and CSC phenotypes and regulation of their functions [273,274].

A large number of studies have provided supportive evidence showing that hypoxia induces the PI3K/Akt/mTOR signaling pathway, consistent with cell proliferation, maintenance of stem cell function via HIF-dependent mechanisms in a wide range of cells including cancer cells [275-281]. The evidence suggests that both HIF and PI3K/Akt/mTOR signaling pathways plays important roles in tumor hypoxic microenvironment, leading to tumor aggressiveness. Thus, there may exist a crosslink between both signaling pathways under hypoxic conditions. Earlier experimental studies have shown that Akt signaling pathway participates in the up-regulation of growth factors-induced HIF-1α activity by increased phosphorylation of elF4E and p70S6K in various cells including cancer cells [282-284]. A number of recent studies have demonstrated that Akt/mTOR up-regulates HIF-1α expression, and stabilizes HIF-1α, and increases HIF-1α synthesis in various cells including cancer cells under normoxic and hypoxic conditions [38,280,281]. It is also known that mTOR stimulates the translation of HIF-1α mRNA into protein [285]. Specific inhibitors for either PI3K, Akt or mTOR decrease the levels of HIF-1α, and inhibit hypoxia-induced HIF-1 transactivation and VEGF expression, leading to a decrease in angiogenesis [260,276, 277,279,285-288]. It has also been shown that the loss of PTEN facilitates HIF-1 mediated gene expression, leading to increased tumor growth and tumor vascularity, compared to PTEN expressing prostate cancer cells [289]. These data suggest that PI3K/Akt/mTOR signaling pathway plays important roles in the regulation of HIF signaling pathway during the development and progression of tumors.

The role of HIF signaling in the regulation of PI3K/Akt/mTOR signaling pathway has been exploited in some studies. One earlier study revealed the inhibition of HIF signaling pathway by 2-methoxyestradiol (2ME), an endogenous HIF-1α inhibitor, which significantly enhances the radiation-induced reduction of cell proliferation and survival, consistent with the down-regulation of radiation-induced expression of HIF-1α and Akt/mTOR signaling, and decreased DNA synthesis in breast cancer cells [290]. These results suggest that HIF signaling pathway participates in the regulation of PI3K/Akt/mTOR signaling pathway, thereby promoting cell proliferation and survival, and tumor growth. Taken together, a crosslink between HIF and PI3K/Akt/mTOR signaling pathways within the tumor microenvironment is highly complex, and thus require in-depth molecular studies in the future.

10. The role of Notch in HIF signaling pathway in tumor

Notch signaling pathway plays an important role in the regulation of a wide range of fundamental cellular processes such as proliferation, stem cell maintenance, differentiation during embryonic and adult development and homeostasis of adult self-renewing organs [291-295]. The proteins encoded by Notch genes (Notch-1, 2, 3, 4) can be activated by interacting with a family of its ligands. The extracellular domain of Notch consists of many EGF-like repeats, which participate in receptor-ligand binding. Notch signaling is initially activated through the receptor-ligand interaction between two neighboring cells. Upon activation, Notch is cleaved, releasing the intracellular domain of the Notch (NICD) through a cascade of proteolytic cleavages by the metalloprotease TNF-α-converting enzyme (TACE) and a γ-secretase complex. The released NICD is then translocated into the nucleus for transcriptional activation of Notch target genes such as Hes1, Hey1, and cyclin D1 [294-296]. The γ-secretase inhibitors (GSI) can suppresses the Notch signaling pathway by binding to γ-secretase complex, leading to the inactivation of this enzyme activity [294,297]. A large number of evidence suggest that Notch signaling pathway plays a pivotal role in tumor aggressiveness by regulation of cell proliferation, apoptosis, invasion, as well as the induction of EMT and CSC functions [291,298-300]. There are some limited reports showing that Notch proteins could exert tumor suppressive effects in few cancers such as lung and skin cancers. However, a large number of studies have demonstrated that Notch proteins have oncogenic in wide range of cancers. Oncogenic or tumor suppressive activities of Notch proteins depend on the cellular context [299].

Several earlier studies have demonstrated that hypoxia-induced HIF-1α binds, stabilizes and activates Notch signaling, leading to cell proliferation and invasion, and maintenance of stem cells in an undifferentiated cell state in various cells including cancer cells as well as stem cells and precursor cells [144,301-303], which have been confirmed by recent findings [304]. Over-expression of Notch-1 can rescue the γ-secretase inhibitor-induced apoptosis of lung cancer cells under hypoxic conditions [301]. It is also known that hypoxia increases Notch-1 mRNA and protein as well as Notch-1 signaling pathway through HIF-1α dependent mechanism [178,301,305,306], supported by recent findings [158,307]. Hypoxic mediated activation of Notch signaling pathway has been shown to involve cell survival, Notch-induced EMT, maintenance of stem cell self-renewal capacity, and tumor cell invasion and migration [178,217,308,309]. It has also been shown that hypoxia induces Notch signaling pathway by binding HIF-1α to Hes-1 promoter regions, leading to the induction of EMT and increased cell invasion and migration in breast cancer [217]. Recently, HIF-2α was found to induce Notch signaling pathway in stem cells [119,120]. Moreover, it has also been documented that the accumulation of both HIF-1α and 2α proteins involves the hypoxic activation of Notch signaling pathway [217]; however, the exact mechanism of how HIF regulates Notch signaling pathway within the tumor microenvironment is not fully understood.

It has been shown that hypoxia induces HIF-mediated Notch-1 signaling via Akt pathway in melanoma cells [305] and that both Notch and HIF-1α are involve in the promotion of hypoxia-induced VEGF expression in mouse embryonic stem cells [310]. Two studies have provided evidence showing that this complex crosstalk may be mediated by FIH. Specifically, Notch proteins contain ankyrin repeat domains which can be hydroxylated by FIH, leading to the reduced Notch signaling pathway [311-313]. It has been speculated that FIH has more binding affinity than HIF-1α, which leads to NICD sequestration of FIH away from HIF-1α, relieving HIF-1α from repression of FIH. Hydroxylation on HIF-1α has also been shown to be decreased in the presence of NICD [311], and therefore, the binding of HIF-1α to HRE promoter sites is more efficient [312]. However, the in-depth investigation is required to elucidate the role and the crosstalk between these two signaling pathways in tumor hypoxic microenvironment.

11. The role of Wnt/β-catenin in HIF signaling pathway in tumor

Wnt/β-catenin signal pathway modulates cell growth by increasing β-catenin levels, β-catenin nuclear localization, and binding to the lymphoid-enhancing factor (LEF)/TCF family of transcription factors, leading to the expression of its target genes in controlling cell growth and proliferation [314-317]. A number of experimental studies have provided evidence for a direct role of Wnt signaling pathway in the embryonic and stem cell development through the regulation of phenotype and self-renewal capacity of stem cells. Moreover, deregulation of Wnt/β-catenin signaling pathway has been observed in many cancers [318,319]. The activation of the Wnt signaling pathway in epidermal stem cells of transgenic mice can result in epithelial cancers [320], and induce EMT phenotype [320-322]. It has also been shown that Wnt/β-catenin signaling pathway is involved in the induction of EMT [174,323]. Therefore, Wnt/β-catenin signaling pathway plays an important role in the regulation of CSCs and EMT [167,174,322,324-329].

Emerging evidence suggests that Wnt/β-catenin pathway plays a role in tumor microenvironment through a crosslink with HIF pathway. The exact mechanism of the crosstalk between these signaling pathways is not fully understood. Two early studies have demonstrated that hypoxia inhibits β-catenin-TCF4 complex formation and transcriptional activity in colon cancer SW480 cell by HIF-1α competition with TCF4 binding directly to β-catenin. However, β-catenin can enhance HIF-mediated transcription, thereby promoting cell survival and adaptation to hypoxia [330,331]. Other multiple studies have shown that hypoxia induces the activation of Wnt/β-catenin pathway by HIF-dependent mechanisms [286,324,332-334]. Furthermore, over-expression of HIF-1α activates Wnt/β-catenin pathway in prostate cancer cells, consistent with cell invasion and the induction of EMT. Conditional knock-down of β-catenin attenuated the EMT characteristics induced by over-expression of HIF-1α [334]. HIF-2α was also shown to enhance the Wnt/β-catenin/TCF signaling pathway by interacting with β-catenin, leading to cell growth and proliferation in renal cancer cells [286]. These data suggest that hypoxic activation of HIF-mediated Wnt/β-catenin pathway, at least in part, contributes to hypoxia-induced EMT in the development and progression of tumor. More interestingly, hypoxia can induce stem cell growth self-renewal, consistent with increased expression of HIF-1α and increased activation of Wnt/β-catenin [335]. Conditional knock-down of HIF-1α inhibits the induction of Wnt effectors, LEF1 and TCF1, potentially by interrupting the HIF-1α binding to the promoters of LEF1 and TCF1 in mouse stem cells, which is in direct agreement with an interesting report [336], suggesting that HIF-mediated Wnt/β-catenin pathway promote the maintenance of stem cell function.

12. The role of Hedgehog in HIF signaling pathway in tumor

Another important signaling pathway potentially associated with the regulation of HIF pathway is the Hedgehog (Hh) signaling pathway. The Hedgehog signaling pathway is a major regulator of cell differentiation, tissue polarity and cell proliferation [337,338]. The Hh family consists of three secreted proteins including sonic Hedgehog (Shh), desert Hedgehog (Dhh), and Indian Hedgehog (Ihh). The Hh proteins are activated by multiple processes such as cleavage, and lipid modification. Binding of lipid-modified Hh to its receptors Patched 1 or Patched 2 (PTCH1 or PTCH2), an inhibitor of Smoothened (Smo) leads to the loss of PTCH activity, and consequent activation of Smo, which transduces the Hh signal to the cytoplasm. Subsequently, the activated Smo causes the activation of Gli (glioma-associated oncogene family zinc finger protein) family of transcriptional effectors through complex interactions with Costal2 (Cos2), Fused (Fu) and Suppressor of fu (Sufu), leading to the up-regulation of PTCH, Wnt, and bone morphogenetic protein (BMP). Therefore, the Hh ligands, Shh, Dhh and Ihh stimulate Gli transcription factors which constitute the final effectors of the Hh signaling pathway.

Emerging evidence clearly suggests the activation of Hh signaling in various human cancers, including basal cell carcinomas, medulloblastomas, leukemia, renal, gastrointestinal, lung, cervical, ovarian, breast and prostate cancers [339,340]. Furthermore, because Hh plays a central role in the control of cell proliferation and differentiation of both embryonic stem cells and adult stem cells, the aberrant activation of Hh signaling could lead to the generation of CSCs and the development of tumor [329]. Recent studies have shown that activation of Hh signaling is critically related to the features of CSCs and EMT in many types of cancers including colon, gastric, esophagus, hepatic, and other cancers [337,338,341,342], suggesting that Hh signaling pathway may play a key role in the regulation of CSCs and EMT within the tumor microenvironment.

Emerging evidence also suggests that Hh signal pathway may be involved in the regulation of HIF pathway under hypoxic conditions. Earlier studies have shown that Hh pathway has a key role in the regulation of angiogenesis [343]. Specifically, hypervascularization of the neuroectoderm has been observed following transgenic overexpression of Shh in the dorsal neural tube of Zebra fish [343]. These observations clearly suggest the role of Hh signal pathway in angiogenesis mediated through the regulation in the expression of pro-angiogenic factors, VEGF, Ang-1 and -2 [344-346]. Thus, there may be a crosstalk between HIF and Hh pathways during cell development and differentiation.

One experimental study reveals that hypoxia induces not only the increased expression of IGF-2 and VEGF but also the activation of their upstream regulatory genes, HIF-1α and Shh in embryonic rat heart. Inactivation of expression of both HIF-1α and Shh pathways by its inhibitors, respectively, resulted in the suppression of each other [347], consistent with recent findings [348].

However, the complex crosstalk between these pathways within the tumor microenvironment remains unclear. Recently, it has been shown that hypoxia activates Hh signaling pathway in a ligand-independent manner by up-regulation of the gene expression of Smo, Gli1, and MMP9 in pancreatic cancer cells. Conditional knockdown of Smo decreases the transcription of Hh downstream targets, Gli1 and MMP9, and suppress tumor invasiveness under hypoxic conditions [349]. One recent clinical study showed that the elevation of Shh expression and low Smo expression are associated with a clinical tumor hypoxic indicator, HP5 (the percentage of oxygen pressure reading in each tumor less than 5 mmHg) in cervical cancer [339]. These data suggest that hypoxia-mediated Hh signaling pathway may play an important role within the tumor microenvironment.

13. The regulatory role of hypoxia-mediated miRNAs in tumor

The microRNAs (miRNAs) are small non-protein-coding RNAs (around 19–24 nucleotides) that function as post-transcriptional gene regulators by its specific binding to the 3′ untranslated region (3′-UTR) of target mRNAs to control protein synthesis or degradation of the mRNAs [350]. miRNAs are currently recognized as regulators of the expression of most genes, and consequently play critical roles in a wide array of biological processes, including cell differentiation, proliferation, death, metabolism and energy homeostasis [314,315]. Sufficient evidence has suggested that miRNAs might have an important role in development and progression of tumors. The altered expression of miRNAs has been associated with clinical prognosis of tumor, resistance to chemo-radiation therapy, tumor recurrence and/or relapse. Tumor hypoxia is a major event during the development and progression of tumor. A large number of miRNAs have been reported to be responsive to hypoxia in a wide range of cells or tissues. Moreover, a number of miRNAs have been identified to be associated with hypoxia and HIF signaling pathways in various cancer cells [351-354]. Thus hypoxia-mediated miRNAs may play an important role in tumor aggressiveness through the regulation of cellular signaling pathways including HIF pathway. Targeting these hypoxia-mediated miRNAs might provide a novel therapeutical strategy for the prevention and/or treatment of cancer. The following paragraphs will provide evidence in support of several miRNAs that are associated with many solid tumors and their regulation under normoxic and hypoxic conditions driving tumor aggressiveness.

13.1. The up-regulation of hypoxia-mediated miRNAs in tumor

13.1.1. The role of miR-21

miR-21 has been considered as a pro-oncogenic molecule. The clinical studies has shown that increased expression of miR-21 is associated with poor clinical prognosis in a wide variety of cancers such as pancreatic cancer, prostate cancer, breast cancer and brain cancer [355,356]. It has been reported that the increased expression of miR-21 results in the decreased expression of PTEN, a known tumor suppressor in cancers [357,358]. The miR-21 has also been reported to show anti-apoptotic, proliferative, invasive and angiogenic properties in cancer cells [356,359,360].

Increased expression ofmiR-21 has been reported in cancer cells such as breast cells under hypoxic conditions [351,352]. Over-expression of miR-21 in DU145 cells increases the expression of HIF-1α and VEGF and increases tumor angiogenesis. Over-expression of miR-21 target gene, PTEN, inhibits tumor angiogenesis through the down-regulation of HIF-1α and VEGF in cancer cells [361]. Thus, HIF-1α is a key downstream target of miR-21 in the regulation of tumor angiogenesis. It has also been shown that over-expression of miR-21 could promote the survival in bone marrow mesenchymal stem cells under hypoxic condition. Down-regulation of miR-21 increase apoptosis of mesenchymal stem cells [362]. Our unpublished data shows that hypoxia leads to increased expression of miR-21 in pancreatic and prostate cancer cells, consistent with the expression of VEGF, and self-renewal capacity of CSC-like cells. The knock-down of miR-21 expression by its siRNA inhibitor decreased the expression of VEGF and self-renewal capacity in pancreatic and prostate cancer sphere-forming cells under hypoxic condition, which suggests that hypoxia-inducedmiR-21 expression might play a critical role within the tumor microenvironment, contributing to tumor aggressiveness.

13.1.2. The role of miR-107

miR-107 has been found to be up-regulated in breast cancer and colon cancer cells under hypoxic condition [351]. A recent study from human colon cancer specimens revealed that miR-107 is a p53 target molecule and p53 directly increases the expression of miR-107 via its 5′-UTR [363]. The over-expression of miR-107 decreases the transactivation of an endogenous HIF-1β 3′-UTR-driven luciferase reporter gene and HIF-1β protein level. Mutation of the miR-107 target sites attenuates this regulation. Moreover, the over-expression of miR-107 decreases the induction of VEGF expression by chemical-induced hypoxic conditions [363]. These data suggest that miR-107 might inhibit VEGF mediated angiogenesis through the down-regulation of HIF-1β-mediated signaling pathway. However, further investigation is required to elucidate the role of miR-107 in tumor aggressiveness and its role within the tumor hypoxic microenvironment.

13.1.3. The contribution of miR-181b

The expression of miR-181b has been reported to be associated with clinical prognosis [364-366]. Altered expression of miR-181b has been found in various malignancies. However, the results are not consistent [364-369]. Several experimental studies have shown that miR-181b increases tumor growth, invasion, metastasis, and multidrug resistance by targeting Bcl-2 and MMP signaling in human cancer cell lines [370-372]. One report indicates that miR-181b function as a tumor suppressor in glioma cells [373]. These data suggest that oncogenic or tumor suppressive activities of miR-181b may depend on the cellular context.

It has been documented that hypoxia induces the expression of miR-181b in various cells including cancer cells [365,374]. The conditional inhibition of miR-181b suppresses proliferation of retinoblastoma cells [375]. The increased expression of miR-181b has been observed in malignant patients with progressive stage, not with stable stage [365]. However, the exact function of hypoxia-mediated miR-181b in tumor microenvironment is still not clear.

13.1.4. The role of miR-210

miR-210 has been identified as hypoxia-induced miRNA in large number of experimental studies. It has been documented that hypoxia induces the increased expression of miR-210 in all the cells tested including cancer cells [351,374,376-383]. Clinically, the increased expression of miR-210 has been associated with poor prognosis in breast cancer and other cancers. The high serum level of miR-210 has also been found in lymphoma patients [384,385]. It has also been shown that the hypoxia-induced miR-210 increases the expression of VEGF and CAIX in pancreatic cancer cells by a HIF-1α-dependent mechanism [378,383], suggesting a regulatory role in tumor angiogenesis [351,353,354,386-388]. Hypoxia-induced miR-210 has also been shown to modulate DNA repair pathway. Forced over-expression of miR-210 was found to suppress the levels of radiation sensitive 52 (RAD52), a key factor in homology-dependent repair (HDR), leading to impaired DNA repair and genetic instability [389]. These data suggest that hypoxia-induced miR-210 plays an important role in the tumor microenvironment, suggesting that targeting miR-210 would provide a novel therapeutic strategy for the treatment of human malignancies.

13.1.5. Contribution of miR-373

miR-373, a hypoxia-mediated miRNA, has been shown to be a mediator of DNA repair systems in various cells including cancer cells via a HRE binding site in the miR-373 promoter [389,390]. It has been shown that miR-373 is up-regulated in a HIF-1α-dependent manner in hypoxic cells [374,389,390]. Moreover, forced over-expression of miR-373 resulted in decreased levels of the nucleotide excision repair proteins, RAD23B and RAD52, both of which are found to be down-regulated in hypoxia, which provides new mechanistic insight into the effect of hypoxia-mediated miRNAs on DNA repair and genetic instability in tumor.

13.2. Down-regulation of hypoxia-mediated miRNAs in tumor

13.2.1. The role of Let-7

A large number of studies have documented that let-7 family has a key regulatory role in the development and progression of tumor by targeting multiple signaling pathways. The expression of let-7 has been shown to be negatively associated with clinical outcomes. Several let-7 family members, for example, let-7b and c have been identified as negative regulators of EMT and CSCs through the regulation of PTEN, CSC marker Lin28b in pancreatic and prostate cancer cells [391-395]. Recently, most of let-7 family members such as let- 7a, b, c, e, f, and g have been reported to be responsive to hypoxic condition in human cancer cells [387]. The results show that hypoxia can lead to the down-regulation of let-7a, b, c, d, e, f, and g in human nasopharyngeal carcinoma cells. One of the hypoxia-mediated let-7 member, let-7b has been validated to be a putative VEGF target miRNA [387], suggesting that let-7 may act a regulatory role in the tumor microenvironment by targeting angiogenesis; however, the exact molecular mechanism of hypoxia-mediated let-7 expression and loss of expression within the tumor hypoxic microenvironment is not clear.

13.2.2. The significance of miR-20

miR-20a and b have been found as hypoxia-mediated miRNAs. Hypoxia results in decreased expression of miR-20a and b in human nasopharyngeal carcinoma cells. The miR-20a and b have been considered as a putative VEGF targeting miRNA [387]. A recent in vitro study has shown that by using pre-miR-20b and anti-miR-20b transfection, miR-20b can inhibit the expression of VEGF in HIF-dependent mechanism under chemically-induced hypoxic condition in breast cancer cells [396], which suggest that miR-20 may be a negative regulator of VEGF activity.

13.2.3. The role of miR-22

Emerging evidence suggests that miR-22 plays an important role in tumorigenesis in a cell type specific manner. It has been shown that miR-22 is up-regulated in human senescent fibroblasts and epithelial cells, but down-regulated in various cancer cell lines such as colon cancer, liver cancer, ovarian cancer, and ER+ breast cancer cells [397-402]. The down-regulation of miR-22 is reported to be associated with poor prognosis of liver cancer [400]. A number of experimental studies have demonstrated that re-expression of miR-22 decreases cell growth, invasion, and metastasis in various cancer cells by targeting PTEN, p21, and p53 [397-400]. These findings suggest that miR-22 may act as a tumor suppressor. It has also been shown that hypoxia causes decreased expression of miR-22 in colon cancer cells. Re-expression of miR-22 results in the inhibition of HIF-1α expression, consistent with repressing the expression of VEGF under hypoxic condition. Knock-down of endogenous miR-22 enhances hypoxia-induced expression of HIF-1α and VEGF in cancer cells, suggesting that hypoxia-mediated miR-22 expression may have an anti-angiogenic effect through the down-regulation of HIF-1α and VEGF expression in the development and progression of tumor [403].

13.2.4. The significance of miR-101

miR-101 has been reported to act as anti-oncogenic molecule. A number of experimental studies have demonstrated that miR-101 plays a protective role in tumor aggressiveness through inhibition of the phenotype and function of CSCs mediated by inactivation of enhancer of zeste homolog 2 (EZH2), an epigenetic regulator of cell survival, proliferation, and CSC function, in various cancer cells [358,404]. The knock-down of EZH2 by its siRNA suppresses cell proliferation, invasion, and the self-renewal capacity of CSCs in human pancreatic cancer MiaPaCa-2 cells, consistent with the inhibition of gene expression of CSCs. Re-expression of miR-101 results in the inhibition of EZH2 expression consistent with the suppression of the self-renewal capacity of CSCs in MiaPaCa-2 cells and its CSC-like sphere cells [358]. It has been reported that hypoxia decreases the expression of miR-101 by HIF-dependent mechanism in prostate cancer cells [405]. Moreover, increased expression of HIF-1α/β by chemically-induced hypoxic condition leads to the down-regulation of miR-101 expression whereas re-expression of miR-101 decreases cancer cell invasion and migration, consistent with decreased expression of EZH2 [405], suggesting hypoxia-mediated expression of miR- 101 may have a key role within the tumor microenvironment via inhibition of CSC function. However, the precise role and the crosstalk between miR-101, HIF and CSC function in tumor aggressiveness require for further in-depth investigation.

The expression of other hypoxia-mediated miRNA, such as 23a,b and miR-26a,b have been reported in several studies [374,387] although the results are not consistent. Altered expression of these miRNAs has also been observed in certain malignancy [406-410]; however, the exact functions of these hypoxia-mediated miRNAs in the development and progression of tumor is required to be defined in the future. It is important to note that targeted inactivation or reexpression of lost miRNAs in the tumor could be a novel therapeutic strategy. Although up-regulation or down-regulation of miRNAs is experimentally possible through gene transfer technology, their application in humans is very limited, suggesting that alternate approach must be devised to regulate miRNAs as discussed in the following paragraphs.

14. Natural product-based HIF inhibitors

In the following paragraphs, we are discussing the role of several natural agents collectively called nutraceutical that may be important regulator of HIF and targeted miRNAs.

14.1. Soy isoflavone

Isoflavones are found primarily in the members of the Leguminosae family. Foods such as soy, lentil, bean, and chickpea are the most common sources of isoflavones; however, soybean is the food that contains abundant amounts of isoflavones. Genistein, daidzein, and glycitein are three major components of isoflavones found in soybeans and soy protein-rich products. Several epidemiological and clinical studies have shown that isoflavone-rich soy products could have protective effects against various cancers such as prostate cancer [411-414].

A large number of experimental studies from our laboratory and others have shown that isoflavones, particularly genistein and daidzein, elicit anti-tumor effects by targeting NF-κB, Wnt, Notch-1, and Akt pathways in many cancers [207,415-420]. Due to its anti-tumor properties, isoflavones have been proposed as potential HIF-1 inhibitor. Our recent studies have shown that soy isoflavone causes sensitization of radiation-induced cell death in prostate cancer cells associated with decreased levels of apurinic/apyrimidinic endonuclease (APE)/redox factor 1 (Ref1), a redox activator of transcription factors, including NF-κB and HIF-1α. These observations suggest that soy isoflavone may show an antitumor effect by targeting HIF-1α pathway. Furthermore, we have demonstrated that pre-treatment with soy isoflavone genistein or daidzein inhibits sarcoma (Src)/signal transducer and activator of transcription 3 (STAT3)/HIF-1α activity induced by radiation and decreased nuclear translocation of HIF-1α [421,422]. These findings suggest that soy isoflavones may have inhibitory effects on HIF-signaling pathway in tumor microenvironment, which could also be due to deregulation of miRNAs. Our recent studies have demonstrated that soy isoflavone treatment increases the expression of let-7b, c, d, and e, and decreases the expression of miR-21 consistent with its anti-tumor activity in human pancreatic cancer cells [394,423,424]. These data suggest that soy isoflavone has an anti-tumor activity, which is in part by targeting HIF pathway and hypoxia-mediated miRNAs in the development and progression of tumor.

14.2. Curcumin

Curcumin (diferuloylmethane) is a bioactive compound, which is derived from the plant Curcuma longa (Linn) grown in tropical Southeast Asia [425-427]. Curcumin has been received considerable attention due to its pronounced anti-inflammatory, anti-oxidative, immunomodulating, anti-atherogenic, and anti-carcinogenic activities. Curcumin has been shown to inhibit the growth of a variety of tumor cells. A large number of experimental and clinical studies have shown that curcumin can induce cell apoptosis, inhibit cell proliferation, cell invasion, and tumor growth, consistent with its regulation of multiple cellular signaling pathways such as NF-κB, Notch, Wnt, and Hh [416,417,420,428-430]. The inhibitory role of curcumin in the regulation of HIF pathway has also been exploited in a number of experimental studies. It has been shown that curcumin has the binding affinity to HIF-1α with the low energy [431], and to decrease the protein levels of HIF-1α, HIF-2α, and HIF-1β, and VEGF in various cancer cells such hepatoma and breast carcinoma cells under hypoxic conditions, consistent with the inhibition of angiogenesis in vitro [432,433]. The xenograft animal study also reveals that curcumin treatment suppresses human liver tumor growth by inhibition of HIF signaling pathway [434].

Recently, we have developed a novel synthetic analog of curcumin, 3,4-difluoro-benzo-curcumin, referred to as difluorinated-curcumin (CDF) [429,435], which shows greater bioavailability in multiple tissues, and also inhibits cell growth, DNA-binding activity of NF-κB, Akt, cyclooxygenase 2 (COX2), and the production of PGE2 and VEGF [357,358]. Moreover, CDF suppresses the tumor growth of human pancreatic cancer in xenograft mouse models, consistent with the inactivation of NF-κB, EZH2, COX2, VEGF, and CSC function [358]. Our unpublished data shows that CDF decreases miR-21 expression, the production of VEGF, and CSC self-renewal capacity, wound healing capacity in pancreatic and prostate cancer cells under hypoxic conditions, consistent with the inhibition of the gene expression of VEGF and HIF-1α, suggesting that CDF could function as a potential inhibitor of HIF signaling pathway in cancer cells. Several recent studies have shown that curcumin or its analog increases the expression of let-7b, c, d, e, and i and miR-22, and decreases the expression of miR-21, consistent with its anti-tumor activity in various cancer cells [357,436-438], suggesting a significant role of curcumin and its analog in anti-tumor activity, which appears to be in part by targeting HIF and hypoxia-mediated miRNAs.

14.3. Tea polyphenols

Consumption of green tea has been associated with human health including the prevention of cancers and heart disease. Several epidemiological studies have shown lower incidence of certain cancers such as prostate cancer among Asian men with a high dietary intake of green tea, which suggests that green tea could be a chemopreventive agent against cancers [439]. One report from Japan’s Public Health Center-based Prospective Study revealed that the high consumption of green tea is associated with decreased risk of advanced prostate cancer [440]. Green tea and its constituents have been investigated both in vitro and in vivo. Green tea contains several catechins including epicatechin (EC), epigallocatechin, epicatechin-3-gallate, and epigallocatechin-3-gallate (EGCG). However, EGCG has been shown to be the most potent for the inhibition of tumorigenesis and the reduction of oxidative stress among these catechins [441]. A number of experimental studies have demonstrated that EGCG exerts an anti-tumor activity in vitro and in vivo, potentially associated with its inhibition of NF-κB, Wnt, and Hh pathways in various cancers, such as lung, colon, prostate, colon, and breast cancers [185,442-448]. The regulatory effect of EGCG on HIF-1α pathway has been reported in several in vitro and in vivo studies although the results are not consistent. Some experimental studies showed the inhibitory effect of EGCG on HIF-1α and VEGF activities in various cancer cells [449-452], consistent with increased expression of let-7 in hepatocellular carcinoma cells [453]. However, other reports have shown some opposite effect of EGCG on HIF through the deregulation of miR-210 and PHD [454-457]. Due to the differences of doses, treatment duration, and cell types, these inconsistent findings require further in-depth investigation to elucidate the role of EGCG in the regulation of HIF pathway and hypoxia-mediated miRNAs in the tumor microenvironment.

14.4. Resveratrol

Resveratrol (trans-3,5,4′-trihydroxystilbene) is a polyphenol compound found in the skin of red grapes, berries and peanuts [458]. Increased evidence from experimental studies has shown that resveratrol could suppress many types of cancers by regulation of cell proliferation and apoptosis mediated through multiple cellular signaling pathways such as Akt, Wnt, Hh, and NF-κB [458-464]. The inhibitory role of resveratrol in the regulation of HIF-pathway has been exploited as documented by increased number of experimental studies. It has been shown that resveratrol exerts a protective effect against the hypoxia-induced lesion in various normal cells by inhibition of HIF-1α activity [465-467]. It has also been shown that resveratrol inhibits the activation of HIF-1α and the expression of VEGF and MMP9, leading to the suppression of hypoxia-mediated cell invasion and migration in various cancer cells [468-470]. Thesefindings suggest that resveratrolmay act as an inhibitor of HIF signal pathway within the tumor microenvironment. There are some limited reports showing that resveratrol treatment could decrease miR-21 expression and increase the expression of miR-22 in lung and colon cancer cells by microarray assay [471,472]. More studies are required to elucidate the significance of resveratrol in the regulation of HIF pathway and hypoxia-mediated miRNAs in tumorigenesis.

14.5. Indole-3-carbinol and 3,3′-diindolylmethane

3,3′-Diindolylmethane (DIM) is one of the dimeric product of indole-3-carbinol (I3C) which is produced from naturally occurring glucosinolates contained in a wide variety of plants including members of the family Cruciferae such as broccoli. Under the acidic conditions of the stomach, I3C undergoes extensive and rapid self-condensation reactions to form several derivatives and DIM is one of the major derivatives. Epidemiological studies have indicated that human exposure to indoles through consumption of cruciferous vegetable could decrease cancer risk [473]. DIM has been shown to reduce oxidative stress and stimulate the expression of anti-oxidant response element-driven gene, suggesting the anti-oxidant function of indole compounds [474,475]. Several experimental studies have shown that DIM inhibits tumorigenesis and cancer cell growth, and induces apoptosis in cancer cells in vitro and in vivo. These findings suggest that DIM could serve as a potent agent for the prevention of tumor progression and/or treatment of cancers, potentially associated with its inhibition of multiple cellular signaling pathways such as NF-κB, Akt, and Wnt as documented by recent studies from our laboratory and others [473-478].

Moreover, one in vitro study has demonstrated that DIM reduces the level of HIF-1α, its transcriptional activity, as well as the gene expression of key hypoxia responsive factors, VEGF, furin, enolase-1, GLUT1, and phosphofructokinase. Its inhibitory effect on HIF activity is consistent with increased rate of the prolyl-hydroxylase- and proteasome-mediated degradation of HIF-1α, thereby leading to decreased rate of HIF-1α transcription [479]. Furthermore, we have recently demonstrated that nutritional grade BR-DIM (absorption-enhanced form) inhibited the radiation-induced activation of NF-κB and HIF-1α DNA activities in prostate cancer cells, thereby leading to enhanced radiation-induced cell killing and tumor growth inhibition [480]. These data clearly suggest that DIM may act as an inhibitory mediator of HIF signaling pathway. Our recent experimental data showed that DIM treatment increases the expression of let- 7b, c, d, and e and decreases the expression of miR-21 and miR-22b in human pancreatic cancer cells [394]. Another report showed increased expression of miR-21 by DIM [481]. Therefore, the regulatory role of DIM in tumorigenesis is in part mediated via targeting HIF pathway and hypoxia-mediated miRNAs; however, the role of DIM still requires further investigation.

14.6. Lycopene

Lycopene is one of the major deep-red pigments responsible for the color of tomatoes and its products. Tomatoes and its products including ketchup, tomato juice, and pizza sauce, are the richest sources of lycopene in the daily diets of people in the USA. Lycopene is a potent anti-oxidant agent. It has been known that lycopene is a biologically active carotenoid displaying high physical quenching rate constant with singlet oxygen, which suggests its high activity as an anti-oxidant agent. One early epidemiological study has demonstrated that frequent consumption of tomato products is associated with a decreased risk of prostate cancer [482]. The inverse associations between plasma lycopene and prostate cancer have also been reported in several human studies [483,484]. Experimental studies also showed that lycopene inhibited cell growth in various cancers through the regulation of cell-cycle-related genes. One in vivo animal study revealed that lycopene has anti-tumor effects that could be potentiated by vitamin E, an anti-oxidant that is also present in tomatoes [485], confirming the anti-tumor activity of lycopene. Our previous phase II clinical trial has shown that lycopene supplements reduce tumor size and prostate-specific antigen (PSA) level in localized prostate cancer [486,487], suggesting its promising effects on prostate cancer prevention and/or treatment. Several experimental studies have shown that its anti-tumor effects are clearly associated with its inhibition of IGF/Akt, Wnt, Wnt/β-catenin, NF-κB, and androgen receptor (AR) signaling [482,484,486-488].

The regulatory effect of lycopene on HIF activity has been exploited recently. One docking analysis study revealed that lycopene has its binding affinity to HIF-1α protein with low energy, suggesting its potential inhibitory activity of HIF-1α [431]. However, in one animal model for cerebral ischemia-reperfusion injury established by middle cerebral artery occlusion (MCAO), the rats pre-treated with the high dose of lycopene (20 mg/kg) for 15 days showed increased expression of HIF-1α mRNA in the cerebral cortex after 24 h of cerebral ischemia-reperfusion injury [489]. To our knowledge, there is no report linking anti-tumor activity of lycopene and hypoxia-mediated miRNAs in the development and progression of tumor. The detailed mechanistic role of lycopene on HIF activity and hypoxia-mediated miRNAs requires further in-depth investigation.

14.7. Vitamin D

Vitamin D is one of the essential nutrients for health in humans. There are two major forms of vitamin D, namely vitamins D2 and D3 in the diets. The active form of vitamin D in the body is 1α,25-dihydroxyvitamin D (1,25-(OH)2D), which can be made from either vitamin D2 or vitamin D3. Several epidemiologic and experimental studies have suggested that higher intakes of vitamin D from food and/or supplements lead to higher blood levels of vitamin D that are associated with reduced risks of cancer [490,491]. It has been reported that reduced levels of active vitamin D in the body resulted in a higher incidence and mortality of prostate cancer [490,491]. These findings suggest that vitamin D could be useful as a preventive agent for cancers. A large number of in vitro and in vivo studies have shown that vitamin D inhibits cell growth, invasion, and angiogenesis in tumors through the regulation of multiple signaling pathways, including Wnt, VEGF, and NF-κB [492,493]. Moreover, one experimental study showed that 1,25-(OH)2D3 (0.1–1 μM) inhibits cell proliferation and protein expression of HIF-1α and VEGF in various human cancer cells under hypoxic conditions. 1,25-(OH)2D3 also inhibited HIF-1 transcriptional activity as well asHIF-1 target genes, VEGF, ET-1, and GLUT1. These inhibitory effects of vitamin D on cancer cells are known to be a HIF-1α-dependent mechanism [492]. However, another in vitro study shows that 1,25-(OH)2D (10 nM) increased both mRNA and protein levels of HIF-1α in H-ras-transfected MCF10A breast epithelial cells and its parental cells under normoxic conditions [494]. Moreover, two recent experimental studies have shown that vitamin D treatment up-regulates the expression of miR-21 and miR-22 and down-regulates the expression of miR-20a and b, andmiR-181b in malignant cells [495,496]. However, the precise role of vitamin D on HIF-signaling and hypoxia-mediated miRNAs is not yet clear. Further investigation is required to clarify the role of vitamin D in the regulation of HIF signaling pathway and hypoxia-mediated miRNAs within the tumor microenvironment.

14.8. Chaetocin

Chaetocin, a class of thiodioxopiperazine, is the naturally occurring agent produced by Chaetomium species fungi and has a wide range of biologic activities, including antimicrobial, anti-inflammatory and anticancer effects [497]. It has been shown that that chaetocin has HIF inhibitor activity in cancer cells through the inhibition of the interaction between HIF-1α and p300, leading to the down-regulation of VEGF expression [498], consistent with recent findings which indicate that chaetocin inhibits the expression of HIF-1α and VEGF in various hepatoma cells, and suppress the growth of HIF-1α+/+ fibrosarcoma in a xenograft mouse model [499]. These findings suggest that chaetocin may act as a potential HIF inhibitor. To our knowledge, there is no report linking chaetocin, HIF pathway, and hypoxia-mediated miRNAs in the tumor microenvironment.

14.9. Manassantins

Manassantins and dineolignans are classes of lignoid-containing naturally occurring compounds which are isolated from the roots of the plant Saururus chinensis and have been known to have various biological activities, such as neuroleptic, anti-inflammatory and acyl-CoA:cholesterol acyl-transferase (ACAT) inhibitory activities and these have been used as supplement [500-502]. The anti-inflammatory activity of manassantins is mediated by its inhibition of NF-κB transactivation activity [503]. Manassantin A and B have been considered as potent HIF inhibitors because manassantin A and B exert their inhibitory activity against the hypoxia-induced HIF-1α and VEGF expression at the concentration range of 1 nM to 1 μM, and a complete inhibition was observed at 100 nM in T47D human breast tumor cells [504,505]. However, the precise molecular mechanism has not been fully elucidated, suggesting further in-depth investigation especially in models with tumor hypoxic microenvironment.

15. Zinc

Zinc is an essential trace mineral element required for human health, and it is known to participate in the activation of approximately 300 enzymes, and is it is also involved in the regulation of over 2000 zinc-dependent transcription factors, which are involved in DNA and protein synthesis, cell division, and other metabolisms. Zinc deficiency can cause growth retardation, delayed sexual maturation, depressed immune response, and cause abnormal cognitive functions. Zinc has been considered as an anti-inflammatory and anti-oxidant agent [506,507]. A large body of epidemiological and clinical studies have demonstrated that zinc deficiency is associated with increased risk of certain cancers, such as prostate, esophageal and oral cancers [508-512]. The data from in vitro and in vivo studies clearly suggest that zinc may have anti-tumor activity mediated through multiple cellular signaling mechanisms including NF-κB [513-516]. One recent study has shown that zinc inhibits hypoxia-induced increase in HIF-1 DNA-binding activity and thereby regulates HIF-1-dependent mRNA expression of EPO, and decreases the stability of HIF-1α in the hypoxic astrocyte cells, consistent with reduction in the nuclear translocation of hypoxia-induced assembly of HIF-1α/HIF-1β complex. These data suggest that zinc may have an important role in the regulation of HIF signaling pathway within the tumor microenvironment [517]; however further pre-clinical and clinical investigation is warranted.

16. Conclusions and perspectives

We attempted to summarize the “state-of-our-knowledge” on the role of hypoxia and HIF signaling pathway in tumor aggressiveness as succinctly as possible and during such attempt we could not cite all the published results, and thus we sincerely apologize to those authors whose work could not be cited. In summary, the evidence has clearly provided a strong support showing that HIF signaling pathway plays an important role in the induction and maintaining of CSC and EMT phenotypes and regulates their functions mediated through the regulation of multiple complex signaling molecules within the tumor microenvironment. Moreover, emerging evidence also suggests that HIF signaling pathway plays an important role in the regulation of NF-κB, PI3K/Akt/mTOR, Notch, Wnt/β-catenin, and Hh signaling not only during normal but also plays important roles during the development and progression of tumors. Furthermore, a number of hypoxia-mediated miRNAs have been shown to mediate cancer cell proliferation, invasion, migration, angiogenesis, and the self-renewal capacity of CSCs through HIF-dependent mechanisms. Therefore, the complexity of the kinship of hypoxia and the regulation of HIF signaling pathway, the biological functions of CSCs, EMT, and their regulation through deregulated expression of miRNAs are important biological processes (Fig. 3), which must be investigated both in pre-clinical and clinical settings. Understanding of precise molecular event regulated by hypoxia would allow in designing targeted approach for the development of newer therapeutics, which would be useful for improving the overall survival of patients diagnosed with malignancies especially solid tumors. To that end, it is our perspective that nutraceutical could serve as promising novel agents to fulfill that requirement, which of course need further preclinical and clinical investigations.

Fig. 3.

Fig. 3

Potential pathways linking hypoxia to CSCs and EMT in tumor aggressiveness. Hypoxic stress increases the stability and expression of the HIFs, especially HIF-1 and 2 in the cells under tumor microenvironment. The activation of HIF signaling pathways up-regulate (1) tumor angiogenesis via the regulation of VEGF, VEGFR1, and Ang-2; (2) the maintenance of CSC phenotype and function via the regulation of Oct4, Nanog, and Sox2; (3) the induction of EMT characteristics via regulation of ZEB2, Snail, Twist, Slug, TGF-β, Wnt, and Notch; and (4) the metabolic adaptation of hypoxic stress via regulation of Glut 1,2, CAIX, LDH-5, IGF-2, and BNIP3. These regulations result in tumor growth, invasion, metastasis, treatment resistance, and recurrence. Complex interlinking of HIF pathway to NF-κB, Akt/mTOR, Notch, Wnt/β-catenin, and Hedgehog signaling pathways increases the contribution of HIF signaling pathway to tumor aggressiveness by the regulation of angiogenesis, CSCs, EMT, and metabolic shifts. Hypoxia also induces the altered expression of miRNAs, for examples, the up-regulation of miR-21, miR-210, and miR-373; and the down-regulation of let-7, miR-20, miR-22, and miR-101 in the tumor cells. The deregulation of the hypoxia-associated miRNAs may enhance tumor angiogenesis, CSC and EMT characteristics through the modulation of multiple signaling pathways. Natural agents such as isoflavones, curcumin, DIM, lycopene, resveratrol, and EGCG have been demonstrated to modulation of these signaling pathways, potentially by the differential regulation of hypoxia-mediated miRNAs, leading to the inhibition/attenuation of tumor aggressiveness.

Acknowledgments

We thank Puschelberg and Guido foundations for their generous financial contribution. We also thank Ms. Ahmedi Bee Fnu, Mr. Anthony Badie Oraha, and Mr. Evan Bao for the technical assistance. Grant support was provided by the National Cancer Institute, NIH grants 5R01CA131151, 3R01CA131151-02S1, 5R01CA132794 and 1R01CA154321 (F.H. Sarkar), DOD Exploration-Hypothesis Development Award PC101482 (B Bao).

List of abbreviations

ABCG2

ATP-binding cassette transporter G2

ACAT

acyl-CoA cholesterol acyl-transferase

Akt

protein kinase B

Ang-2

angiopoietin 2

APE

apurinic/apyrimidinic endonuclease

AR

androgen receptor

Bcl2

B-cell lymphoma 2

bHLH

basic helix–loop–helix

bHLH-PAS

Per-Arnt/AhR-Sim basic helix–loop–helix

BMP

bone morphogenetic protein

BNIP3

Bcl2/adenovirus E1B 19 kDa protein-interacting protein 3

BR-DIM

absorption-enhanced form of DIM

CAD

COOH-terminal transactivation domain

CAIX

carbonic anhydrase 9

CBP

CREB binding protein

CD44

the cluster of differentiation 44

CDF

difluorinated-curcumin

CDKN1A

cyclin-dependent kinase inhibitor 1A

c-MET

mesenchymal–epithelial transition factor

Cos2

Costal2

COX2

cyclooxygenase 2

CSC/CSCs

cancer stem cell/cancer stem cells

CXCR4

C–X–C chemokine receptor type 4

Dhh

desert Hedgehog

DIM

3,3′-diindolylmethane

EC

epicatechin

EGCG

epigallocatechin-3-gallate

elF4E

eukaryotic initiation factor 4E

EMT

epithelial-to-mesenchymal transition

EPO

erythropoietin

ET-1

endothelin 1

EZH2

enhancer of zeste homolog 2

FIH

the factor-inhibiting HIF

Flt-1

fms-like tyrosine kinase 1

Fu

fused

Gli

glioma-associated oncogene family zinc finger protein

GLUT

glucose transporters

GSI

γ-secretase inhibitors

HAF

hypoxia associated factor

Hey1

hairy/enhancer-of-split related with YRPW motif 1

HDCA3

histone deacelylase 3

HDR

homology-dependent repair

Hh

Hedgehog

HRE

hypoxia-response element

Hsp90

heat shock protein 90

I3C

indole-3-carbinol

IGF-1

insulin-like growth factor 1

Ihh

Indian Hedgehog

IL-6

interleukin 6

IκB

inhibitory proteins of NF-κB

IKK

IκB kinase

iNOS

inducible nitric oxide synthase

IPAS

inhibitory PAS protein

LDH5

lactate dehydrogenase 5

LEF

lymphoid-enhancing factor

LEF

lymphoid-enhancing factor

Lin-28

a conserved regulator of cell fate succession in animals

LKB1

liver kinase B1

LOX

lysyl oxidase

LPS

liposaccharides

miRNAs

microRNAs

2ME

2-methoxyestradiol

MMP

matrix metalloproteinases

mTOR

mammalian target of rapamycin

mTORC

mTOR complex

NF-κB

nuclear factor of κB

NICD

intracellular domain of the Notch

Notch1

Notch homolog 1

NSCLC

non-small cell lung cancer

Oct4

octamer-binding transcription factor 4

ODD

oxygen-dependent degradation domain

P70S6K1

p70-S6 kinase 1

PAS

per-aryl hydrocarbon receptor nuclear translocator (ARNT)-sim

PDK1

pyruvate dehydrogenase kinase 1

PHD

prolyl hydroxylase

PSA

prostate-specific antigen

PTEN

phosphatase and tensin homolog deleted on chromosome 10

RAD52

radiation sensitive 52, a DNA damage repair protein

Ref1

redox factor 1

ROS

reactive oxygen species

S6K1

ribosomal protein S6 kinase

SCID

severe combined immunodeficiency

Shh

sonic Hedgehog

SIP2

septin interacting protein 2

Smo

smoothened

Sox2

sex determining region Y box 2

STAT

signal transducer and activator of transcription

Sufu

suppressor of fu

TACE

TNF-α-converting enzyme

TGF-α

transforming growth factor-α

TAK1

TGF-β activated kinase 1

TCF

T cell factors

TIC

tumor initiating cells

Tie2

tyrosine kinase receptor with immunoglobulin and epidermal growth factor homology 2

TNF-α

tumor necrosis factor α

TSC

tuberous sclerosis

uPA

urokinase-type plasminogen activator

3′-UTR

3′-untranslated region

VEGF

vascular endothelial growth factor

VEGFR1

VEGF receptor 1

VHL

von Hippel–Lindau

XIAP

X-linked inhibitor of apoptosis protein

ZEB1

zinc-finger E-box binding homeobox 1

ZO-1

zonula occludens-1

References

  • 1.Fyles A, Milosevic M, Hedley D, Pintilie M, Levin W, Manchul L, Hill RP. Tumor hypoxia has independent predictor impact only in patients with node-negative cervix cancer. J Clin Oncol. 2002;20:680–687. doi: 10.1200/JCO.2002.20.3.680. [DOI] [PubMed] [Google Scholar]
  • 2.Hockel M, Schlenger K, Aral B, Mitze M, Schaffer U, Vaupel P. Association between tumor hypoxia and malignant progression in advanced cancer of the uterine cervix. Cancer Res. 1996;56:4509–4515. [PubMed] [Google Scholar]
  • 3.Jubb AM, Buffa FM, Harris AL. Assessment of tumour hypoxia for prediction of response to therapy and cancer prognosis. J Cell Mol Med. 2010;14:18–29. doi: 10.1111/j.1582-4934.2009.00944.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Moulder JE, Rockwell S. Tumor hypoxia: its impact on cancer therapy. Cancer Metastasis Rev. 1987;5:313–341. doi: 10.1007/BF00055376. [DOI] [PubMed] [Google Scholar]
  • 5.Teicher BA. Hypoxia and drug resistance. Cancer Metastasis Rev. 1994;13:139–168. doi: 10.1007/BF00689633. [DOI] [PubMed] [Google Scholar]
  • 6.Favaro E, Lord S, Harris AL, Buffa FM. Gene expression and hypoxia in breast cancer. Genome Med. 2011;3:55. doi: 10.1186/gm271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Denko NC. Hypoxia, HIF1 and glucose metabolism in the solid tumour. Nat Rev Cancer. 2008;8:705–713. doi: 10.1038/nrc2468. [DOI] [PubMed] [Google Scholar]
  • 8.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]
  • 9.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–C1180. doi: 10.1152/ajpcell.1996.271.4.C1172. [DOI] [PubMed] [Google Scholar]
  • 10.Semenza GL. HIF-1 and tumor progression: pathophysiology and therapeutics. Trends Mol Med. 2002;8:S62–S67. doi: 10.1016/s1471-4914(02)02317-1. [DOI] [PubMed] [Google Scholar]
  • 11.Semenza GL. Involvement of hypoxia-inducible factor 1 in human cancer. Intern Med. 2002;41:79–83. doi: 10.2169/internalmedicine.41.79. [DOI] [PubMed] [Google Scholar]
  • 12.Semenza GL. Targeting HIF-1 for cancer therapy. Nat Rev Cancer. 2003;3:721–732. doi: 10.1038/nrc1187. [DOI] [PubMed] [Google Scholar]
  • 13.Semenza GL. Defining the role of hypoxia-inducible factor 1 in cancer biology and therapeutics. Oncogene. 2010;29:625–634. doi: 10.1038/onc.2009.441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Pugh CW, Ratcliffe PJ. Regulation of angiogenesis by hypoxia: role of the HIF system. Nat Med. 2003;9:677–684. doi: 10.1038/nm0603-677. [DOI] [PubMed] [Google Scholar]
  • 15.Zhou J, Schmid T, Schnitzer S, Brune B. Tumor hypoxia and cancer progression. Cancer Lett. 2006;237:10–21. doi: 10.1016/j.canlet.2005.05.028. [DOI] [PubMed] [Google Scholar]
  • 16.Eppenberger U, Kueng W, Schlaeppi JM, Roesel JL, Benz C, Mueller H, Matter A, Zuber M, Luescher K, Litschgi M, Schmitt M, Foekens JA, Eppenberger-Castori S. Markers of tumor angiogenesis and proteolysis independently define high- and low-risk subsets of node-negative breast cancer patients. J Clin Oncol. 1998;16:3129–3136. doi: 10.1200/JCO.1998.16.9.3129. [DOI] [PubMed] [Google Scholar]
  • 17.Gasparini G, Toi M, Gion M, Verderio P, Dittadi R, Hanatani M, Matsubara I, Vinante O, Bonoldi E, Boracchi P, Gatti C, Suzuki H, Tominaga T. Prognostic significance of vascular endothelial growth factor protein in node-negative breast carcinoma. J Natl Cancer Inst. 1997;89:139–147. doi: 10.1093/jnci/89.2.139. [DOI] [PubMed] [Google Scholar]
  • 18.Linderholm B, Grankvist K, Wilking N, Johansson M, Tavelin B, Henriksson R. Correlation of vascular endothelial growth factor content with recurrences, survival, and first relapse site in primary node-positive breast carcinoma after adjuvant treatment. J Clin Oncol. 2000;18:1423–1431. doi: 10.1200/JCO.2000.18.7.1423. [DOI] [PubMed] [Google Scholar]
  • 19.Wong C, Wellman TL, Lounsbury KM. VEGF and HIF-1alpha expression are increased in advanced stages of epithelial ovarian cancer. Gynecol Oncol. 2003;91:513–517. doi: 10.1016/j.ygyno.2003.08.022. [DOI] [PubMed] [Google Scholar]
  • 20.Brennan DJ, Jirstrom K, Kronblad A, Millikan RC, Landberg G, Duffy MJ, Ryden L, Gallagher WM, O’Brien SL. CA IX is an independent prognostic marker in premenopausal breast cancer patients with one to three positive lymph nodes and a putative marker of radiation resistance. Clin Cancer Res. 2006;12:6421–6431. doi: 10.1158/1078-0432.CCR-06-0480. [DOI] [PubMed] [Google Scholar]
  • 21.Chia SK, Wykoff CC, Watson PH, Han C, Leek RD, Pastorek J, Gatter KC, Ratcliffe P, Harris AL. Prognostic significance of a novel hypoxia-regulated marker, carbonic anhydrase IX, in invasive breast carcinoma. J Clin Oncol. 2001;19:3660–3668. doi: 10.1200/JCO.2001.19.16.3660. [DOI] [PubMed] [Google Scholar]
  • 22.Generali D, Fox SB, Berruti A, Brizzi MP, Campo L, Bonardi S, Wigfield SM, Bruzzi P, Bersiga A, Allevi G, Milani M, Aguggini S, Dogliotti L, Bottini A, Harris AL. Role of carbonic anhydrase IX expression in prediction of the efficacy and outcome of primary epirubicin/tamoxifen therapy for breast cancer. Endocr Relat Cancer. 2006;13:921–930. doi: 10.1677/erc.1.01216. [DOI] [PubMed] [Google Scholar]
  • 23.Generali D, Berruti A, Brizzi MP, Campo L, Bonardi S, Wigfield S, Bersiga A, Allevi G, Milani M, Aguggini S, Gandolfi V, Dogliotti L, Bottini A, Harris AL, Fox SB. Hypoxia-inducible factor-1alpha expression predicts a poor response to primary chemoendocrine therapy and disease-free survival in primary human breast cancer. Clin Cancer Res. 2006;12:4562–4568. doi: 10.1158/1078-0432.CCR-05-2690. [DOI] [PubMed] [Google Scholar]
  • 24.Hussain SA, Ganesan R, Reynolds G, Gross L, Stevens A, Pastorek J, Murray PG, Perunovic B, Anwar MS, Billingham L, James ND, Spooner D, Poole CJ, Rea DW, Palmer DH. Hypoxia-regulated carbonic anhydrase IX expression is associated with poor survival in patients with invasive breast cancer. Br J Cancer. 2007;96:104–109. doi: 10.1038/sj.bjc.6603530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Trastour C, Benizri E, Ettore F, Ramaioli A, Chamorey E, Pouyssegur J, Berra E. HIF-1alpha and CA IX staining in invasive breast carcinomas: prognosis and treatment outcome. Int J Cancer. 2007;120:1451–1458. doi: 10.1002/ijc.22436. [DOI] [PubMed] [Google Scholar]
  • 26.Van den Eynden GG, Van der Auwera I, Van Laere SJ, Colpaert CG, Turley H, Harris AL, van DP, Dirix LY, Vermeulen PB, Van Marck EA. Angiogenesis and hypoxia in lymph node metastases is predicted by the angiogenesis and hypoxia in the primary tumour in patients with breast cancer. Br J Cancer. 2005;93:1128–1136. doi: 10.1038/sj.bjc.6602828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Brahimi-Horn MC, Bellot G, Pouyssegur J. Hypoxia and energetic tumour metabolism. Curr Opin Genet Dev. 2011;21:67–72. doi: 10.1016/j.gde.2010.10.006. [DOI] [PubMed] [Google Scholar]
  • 28.Giatromanolaki A, Sivridis E, Gatter KC, Turley H, Harris AL, Koukourakis MI. Lactate dehydrogenase 5 (LDH-5) expression in endometrial cancer relates to the activated VEGF/VEGFR2(KDR) pathway and prognosis. Gynecol Oncol. 2006;103:912–918. doi: 10.1016/j.ygyno.2006.05.043. [DOI] [PubMed] [Google Scholar]
  • 29.Kang SS, Chun YK, Hur MH, Lee HK, Kim YJ, Hong SR, Lee JH, Lee SG, Park YK. Clinical significance of glucose transporter 1 (GLUT1) expression in human breast carcinoma. Jpn J Cancer Res. 2002;93:1123–1128. doi: 10.1111/j.1349-7006.2002.tb01214.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Koukourakis MI, Giatromanolaki A, Sivridis E, Simopoulos C, Turley H, Talks K, Gatter KC, Harris 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;53:1192–1202. doi: 10.1016/s0360-3016(02)02848-1. [DOI] [PubMed] [Google Scholar]
  • 31.Koukourakis MI, Giatromanolaki A, Sivridis E, Bougioukas G, Didilis V, Gatter KC, Harris AL. Lactate dehydrogenase-5 (LDH-5) overexpression in non-small-cell lung cancer tissues is linked to tumour hypoxia, angiogenic factor production and poor prognosis. Br J Cancer. 2003;89:877–885. doi: 10.1038/sj.bjc.6601205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Koukourakis MI, Giatromanolaki A, Simopoulos C, Polychronidis A, Sivridis E. Lactate dehydrogenase 5 (LDH5) relates to up-regulated hypoxia inducible factor pathway and metastasis in colorectal cancer. Clin Exp Metastasis. 2005;22:25–30. doi: 10.1007/s10585-005-2343-7. [DOI] [PubMed] [Google Scholar]
  • 33.Koukourakis MI, Giatromanolaki A, Sivridis E, Gatter KC, Harris AL. Lactate dehydrogenase 5 expression in operable colorectal cancer: strong association with survival and activated vascular endothelial growth factor pathway—a report of the Tumour Angiogenesis Research Group. J Clin Oncol. 2006;24:4301–4308. doi: 10.1200/JCO.2006.05.9501. [DOI] [PubMed] [Google Scholar]
  • 34.Koukourakis MI, Giatromanolaki A, Winter S, Leek R, Sivridis E, Harris AL. Lactate dehydrogenase 5 expression in squamous cell head and neck cancer relates to prognosis following radical or postoperative radiotherapy. Oncology. 2009;77:285–292. doi: 10.1159/000259260. [DOI] [PubMed] [Google Scholar]
  • 35.Giatromanolaki A, Koukourakis MI, Sowter HM, Sivridis E, Gibson S, Gatter KC, Harris AL. BNIP3 expression is linked with hypoxia-regulated protein expression and with poor prognosis in non-small cell lung cancer. Clin Cancer Res. 2004;10:5566–5571. doi: 10.1158/1078-0432.CCR-04-0076. [DOI] [PubMed] [Google Scholar]
  • 36.Tan EY, Campo L, Han C, Turley H, Pezzella F, Gatter KC, Harris AL, Fox SB. BNIP3 as a progression marker in primary human breast cancer; opposing functions in in situ versus invasive cancer. Clin Cancer Res. 2007;13:467–474. doi: 10.1158/1078-0432.CCR-06-1466. [DOI] [PubMed] [Google Scholar]
  • 37.Giatromanolaki A, Koukourakis MI, Sowter HM, Sivridis E, Gibson S, Gatter KC, Harris AL. BNIP3 expression is linked with hypoxia-regulated protein expression and with poor prognosis in non-small cell lung cancer. Clin Cancer Res. 2004;10:5566–5571. doi: 10.1158/1078-0432.CCR-04-0076. [DOI] [PubMed] [Google Scholar]
  • 38.Cassavaugh JM, Hale SA, Wellman TL, Howe AK, Wong C, Lounsbury KM. Negative regulation of HIF-1alpha by an FBW7-mediated degradation pathway during hypoxia. J Cell Biochem. 2011;112:3882–3890. doi: 10.1002/jcb.23321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Chen C, Ma Q, Ma X, Liu Z, Liu X. Association of elevated HIF-2alpha levels with low Beclin 1 expression and poor prognosis in patients with chondrosarcoma. Ann Surg Oncol. 2011;18:2364–2372. doi: 10.1245/s10434-011-1587-5. [DOI] [PubMed] [Google Scholar]
  • 40.Gong K, Zhang N, Zhang K, Na Y. The relationship of erythropoietin overexpression with von Hippel–Lindau tumour suppressor gene mutations between hypoxia-inducible factor-1alpha and -2alpha in sporadic clear cell renal carcinoma. Int J Mol Med. 2010;26:907–912. doi: 10.3892/ijmm_00000541. [DOI] [PubMed] [Google Scholar]
  • 41.Kawanaka T, Kubo A, Ikushima H, Sano T, Takegawa Y, Nishitani H. Prognostic significance of HIF-2alpha expression on tumor infiltrating macrophages in patients with uterine cervical cancer undergoing radiotherapy. J Med Invest. 2008;55:78–86. doi: 10.2152/jmi.55.78. [DOI] [PubMed] [Google Scholar]
  • 42.Wei L, Liu X, Hu C. Correlation between expression of HIF-2alpha and OCT-4 and prognosis of NSCLC. Zhong Nan Da Xue Xue Bao Yi Xue Ban. 2011;36:854–858. doi: 10.3969/j.issn.1672-7347.2011.09.007. [DOI] [PubMed] [Google Scholar]
  • 43.Foekens JA, Peters HA, Grebenchtchikov N, Look MP, Meijer-van Gelder ME, Geurts-Moespot A, van der Kwast TH, Sweep CG, Klijn JG. High tumor levels of vascular endothelial growth factor predict poor response to systemic therapy in advanced breast cancer. Cancer Res. 2001;61:5407–5414. [PubMed] [Google Scholar]
  • 44.Span PN, Bussink J, Manders P, Beex LV, Sweep CG. Carbonic anhydrase-9 expression levels and prognosis in human breast cancer: association with treatment outcome. Br J Cancer. 2003;89:271–276. doi: 10.1038/sj.bjc.6601122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Unruh A, Ressel A, Mohamed HG, Johnson RS, Nadrowitz R, Richter E, Katschinski DM, Wenger RH. The hypoxia-inducible factor-1 alpha is a negative factor for tumor therapy. Oncogene. 2003;22:3213–3220. doi: 10.1038/sj.onc.1206385. [DOI] [PubMed] [Google Scholar]
  • 46.Carmeliet P, Jain RK. Principles and mechanisms of vessel normalization for cancer and other angiogenic diseases. Nat Rev Drug Discov. 2011;10:417–427. doi: 10.1038/nrd3455. [DOI] [PubMed] [Google Scholar]
  • 47.Generali D, Buffa FM, Berruti A, Brizzi MP, Campo L, Bonardi S, Bersiga A, Allevi G, Milani M, Aguggini S, Papotti M, Dogliotti L, Bottini A, Harris AL, Fox SB. Phosphorylated ERalpha, HIF-1alpha, and MAPK signaling as predictors of primary endocrine treatment response and resistance in patients with breast cancer. J Clin Oncol. 2009;27:227–234. doi: 10.1200/JCO.2007.13.7083. [DOI] [PubMed] [Google Scholar]
  • 48.Mehta S, Hughes NP, Buffa FM, Li SP, Adams RF, Adwani A, Taylor NJ, Levitt NC, Padhani AR, Makris A, Harris AL. Assessing early therapeutic response to bevacizumab in primary breast cancer using magnetic resonance imaging and gene expression profiles. J Natl Cancer Inst Monogr. 2011;2011:71–74. doi: 10.1093/jncimonographs/lgr027. [DOI] [PubMed] [Google Scholar]
  • 49.Rapisarda A, Hollingshead M, Uranchimeg B, Bonomi CA, Borgel SD, Carter JP, Gehrs B, Raffeld M, Kinders RJ, Parchment R, Anver MR, Shoemaker RH, Melillo G. Increased antitumor activity of bevacizumab in combination with hypoxia inducible factor-1 inhibition. Mol Cancer Ther. 2009;8:1867–1877. doi: 10.1158/1535-7163.MCT-09-0274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Sathornsumetee S, Cao Y, Marcello JE, Herndon JE, McLendon RE, Desjardins A, Friedman HS, Dewhirst MW, Vredenburgh JJ, Rich JN. Tumor angiogenic and hypoxic profiles predict radiographic response and survival in malignant astrocytoma patients treated with bevacizumab and irinotecan. J Clin Oncol. 2008;26:271–278. doi: 10.1200/JCO.2007.13.3652. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Overgaard J. Hypoxic modification of radiotherapy in squamous cell carcinoma of the head and neck—a systematic review and meta-analysis. Radiother Oncol. 2011;100:22–32. doi: 10.1016/j.radonc.2011.03.004. [DOI] [PubMed] [Google Scholar]
  • 52.Nordsmark M, Bentzen SM, Rudat V, Brizel D, Lartigau E, Stadler P, Becker A, Adam M, Molls M, Dunst J, Terris DJ, Overgaard J. Prognostic value of tumor oxygenation in 397 head and neck tumors after primary radiation therapy. An international multi-center study. Radiother Oncol. 2005;77:18–24. doi: 10.1016/j.radonc.2005.06.038. [DOI] [PubMed] [Google Scholar]
  • 53.Strofer M, Jelkmann W, Metzen E, Brockmeier U, Dunst J, Depping R. Stabilisation and knockdown of HIF—two distinct ways comparably important in radiotherapy. Cell Physiol Biochem. 2011;28:805–812. doi: 10.1159/000335794. [DOI] [PubMed] [Google Scholar]
  • 54.Liao D, Johnson RS. Hypoxia: a key regulator of angiogenesis in cancer. Cancer Metastasis Rev. 2007;26:281–290. doi: 10.1007/s10555-007-9066-y. [DOI] [PubMed] [Google Scholar]
  • 55.Aebersold DM, Burri P, Beer KT, Laissue J, Djonov V, Greiner RH, Semenza GL. Expression of hypoxia-inducible factor-1alpha: a novel predictive and prognostic parameter in the radiotherapy of oropharyngeal cancer. Cancer Res. 2001;61:2911–2916. [PubMed] [Google Scholar]
  • 56.Hockel M, Vaupel P. Tumor hypoxia: definitions and current clinical, biologic, and molecular aspects. J Natl Cancer Inst. 2001;93:266–276. doi: 10.1093/jnci/93.4.266. [DOI] [PubMed] [Google Scholar]
  • 57.Piret JP, Mottet D, Raes M, Michiels C. CoCl2, a chemical inducer of hypoxia-inducible factor-1, and hypoxia reduce apoptotic cell death in hepatoma cell line HepG2. Ann N Y Acad Sci. 2002;973:443–447. doi: 10.1111/j.1749-6632.2002.tb04680.x. [DOI] [PubMed] [Google Scholar]
  • 58.Wang JH, Wu QD, Bouchier-Hayes D, Redmond HP. Hypoxia upregulates Bcl-2 expression and suppresses interferon-gamma induced antiangiogenic activity in human tumor derived endothelial cells. Cancer. 2002;94:2745–2755. doi: 10.1002/cncr.10520. [DOI] [PubMed] [Google Scholar]
  • 59.Qiao Q, Nozaki Y, Sakoe K, Komatsu N, Kirito K. NF-kappaB mediates aberrant activation of HIF-1 in malignant lymphoma. Exp Hematol. 2010;38:1199–1208. doi: 10.1016/j.exphem.2010.08.007. [DOI] [PubMed] [Google Scholar]
  • 60.Harada H. How can we overcome tumor hypoxia in radiation therapy? J Radiat Res (Tokyo) 2011;52:545–556. doi: 10.1269/jrr.11056. [DOI] [PubMed] [Google Scholar]
  • 61.Wilson WR, Hay MP. Targeting hypoxia in cancer therapy. Nat Rev Cancer. 2011;11:393–410. doi: 10.1038/nrc3064. [DOI] [PubMed] [Google Scholar]
  • 62.da S, Rutter XJ, Rutter GA. Involvement of Per-Arnt-Sim (PAS) kinase in the stimulation of preproinsulin and pancreatic duodenum homeobox 1 gene expression by glucose. Proc Natl Acad Sci U S A. 2004;101:8319–8324. doi: 10.1073/pnas.0307737101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.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]
  • 64.Semenza GL, Jiang BH, Leung SW, Passantino R, Concordet JP, Maire P, Giallongo A. Hypoxia response elements in the aldolase A, enolase 1, and lactate dehydrogenase A gene promoters contain essential binding sites for hypoxia-inducible factor 1. J Biol Chem. 1996;271:32529–32537. doi: 10.1074/jbc.271.51.32529. [DOI] [PubMed] [Google Scholar]
  • 65.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]
  • 66.Wang GL, Jiang BH, Rue EA, Semenza GL. Hypoxia-inducible factor 1 is a basic-helix–loop–helix-PAS heterodimer regulated by cellular O2 tension. Proc Natl Acad Sci U S A. 1995;92:5510–5514. doi: 10.1073/pnas.92.12.5510. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Wang GL, Semenza GL. General involvement of hypoxia-inducible factor 1 in transcriptional response to hypoxia. Proc Natl Acad Sci U S A. 1993;90:4304–4308. doi: 10.1073/pnas.90.9.4304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Ratcliffe PJ, O’Rourke JF, Maxwell PH, Pugh CW. Oxygen sensing, hypoxia-inducible factor-1 and the regulation of mammalian gene expression. J Exp Biol. 1998;201:1153–1162. doi: 10.1242/jeb.201.8.1153. [DOI] [PubMed] [Google Scholar]
  • 69.Ratcliffe PJ, Pugh CW, Maxwell PH. Targeting tumors through the HIF system. Nat Med. 2000;6:1315–1316. doi: 10.1038/82113. [DOI] [PubMed] [Google Scholar]
  • 70.Semenza GL. Hypoxia-inducible factor 1: master regulator of O2 homeostasis. Curr Opin Genet Dev. 1998;8:588–594. doi: 10.1016/s0959-437x(98)80016-6. [DOI] [PubMed] [Google Scholar]
  • 71.Keith B, Adelman DM, Simon MC. Targeted mutation of the murine arylhydrocarbon receptor nuclear translocator 2 (Arnt2) gene reveals partial redundancy with Arnt. Proc Natl Acad Sci U S A. 2001;98:6692–6697. doi: 10.1073/pnas.121494298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Keith B, Simon MC. Hypoxia-inducible factors, stem cells, and cancer. Cell. 2007;129:465–472. doi: 10.1016/j.cell.2007.04.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Maltepe E, Schmidt JV, Baunoch D, Bradfield CA, Simon MC. Abnormal angiogenesis and responses to glucose and oxygen deprivation in mice lacking the protein ARNT. Nature. 1997;386:403–407. doi: 10.1038/386403a0. [DOI] [PubMed] [Google Scholar]
  • 74.Jiang BH, Agani F, Passaniti A, Semenza GL. V-SRC induces expression of hypoxia-inducible factor 1 (HIF-1) and transcription of genes encoding vascular endothelial growth factor and enolase 1: involvement of HIF-1 in tumor progression. Cancer Res. 1997;57:5328–5335. [PubMed] [Google Scholar]
  • 75.Maxwell PH, Dachs GU, Gleadle JM, Nicholls LG, Harris AL, Stratford IJ, Hankinson O, Pugh CW, Ratcliffe PJ. Hypoxia-inducible factor-1 modulates gene expression in solid tumors and influences both angiogenesis and tumor growth. Proc Natl Acad Sci U S A. 1997;94:8104–8109. doi: 10.1073/pnas.94.15.8104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Huang LE, Arany Z, Livingston DM, Bunn HF. Activation of hypoxia-inducible transcription factor depends primarily upon redox-sensitive stabilization of its alpha subunit. J Biol Chem. 1996;271:32253–32259. doi: 10.1074/jbc.271.50.32253. [DOI] [PubMed] [Google Scholar]
  • 77.Jaakkola P, Mole DR, Tian YM, Wilson MI, Gielbert J, Gaskell SJ, Kriegsheim A, Hebestreit HF, Mukherji M, Schofield CJ, Maxwell PH, Pugh CW, Ratcliffe PJ. Targeting of HIF-alpha to the von Hippel–Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science. 2001;292:468–472. doi: 10.1126/science.1059796. [DOI] [PubMed] [Google Scholar]
  • 78.Lando D, Peet DJ, Gorman JJ, Whelan DA, Whitelaw ML, Bruick RK. FIH-1 is an asparaginyl hydroxylase enzyme that regulates the transcriptional activity of hypoxia-inducible factor. Genes Dev. 2002;16:1466–1471. doi: 10.1101/gad.991402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Lisy K, Peet DJ. Turn me on: regulating HIF transcriptional activity. Cell Death Differ. 2008;15:642–649. doi: 10.1038/sj.cdd.4402315. [DOI] [PubMed] [Google Scholar]
  • 80.Pouyssegur J, Dayan F, Mazure NM. Hypoxia signalling in cancer and approaches to enforce tumour regression. Nature. 2006;441:437–443. doi: 10.1038/nature04871. [DOI] [PubMed] [Google Scholar]
  • 81.Iwai K, Yamanaka K, Kamura T, Minato N, Conaway RC, Conaway JW, Klausner RD, Pause A. Identification of the von Hippel–lindau tumor-suppressor protein as part of an active E3 ubiquitin ligase complex. Proc Natl Acad Sci U S A. 1999;96:12436–12441. doi: 10.1073/pnas.96.22.12436. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Maxwell PH, Wiesener MS, Chang GW, Clifford SC, Vaux EC, Cockman ME, Wykoff CC, Pugh CW, Maher ER, Ratcliffe PJ. The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature. 1999;399:271–275. doi: 10.1038/20459. [DOI] [PubMed] [Google Scholar]
  • 83.Ohh M, Park CW, Ivan M, Hoffman MA, Kim TY, Huang LE, Pavletich N, Chau V, Kaelin WG. Ubiquitination of hypoxia-inducible factor requires direct binding to the beta-domain of the von Hippel–Lindau protein. Nat Cell Biol. 2000;2:423–427. doi: 10.1038/35017054. [DOI] [PubMed] [Google Scholar]
  • 84.Yu F, White SB, Zhao Q, Lee FS. HIF-1alpha binding to VHL is regulated by stimulus-sensitive proline hydroxylation. Proc Natl Acad Sci U S A. 2001;98:9630–9635. doi: 10.1073/pnas.181341498. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Berra E, Richard DE, Gothie E, Pouyssegur J. HIF-1-dependent transcriptional activity is required for oxygen-mediated HIF-1alpha degradation. FEBS Lett. 2001;491:85–90. doi: 10.1016/s0014-5793(01)02159-7. [DOI] [PubMed] [Google Scholar]
  • 86.Cockman ME, Masson N, Mole DR, Jaakkola P, Chang GW, Clifford SC, Maher ER, Pugh CW, Ratcliffe PJ, Maxwell PH. Hypoxia inducible factor-alpha binding and ubiquitylation by the von Hippel–Lindau tumor suppressor protein. J Biol Chem. 2000;275:25733–25741. doi: 10.1074/jbc.M002740200. [DOI] [PubMed] [Google Scholar]
  • 87.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;274:6519–6525. doi: 10.1074/jbc.274.10.6519. [DOI] [PubMed] [Google Scholar]
  • 88.Horii K, Suzuki Y, Kondo Y, Akimoto M, Nishimura T, Yamabe Y, Sakaue M, Sano T, Kitagawa T, Himeno S, Imura N, Hara S. Androgen-dependent gene expression of prostate-specific antigen is enhanced synergistically by hypoxia in human prostate cancer cells. Mol Cancer Res. 2007;5:383–391. doi: 10.1158/1541-7786.MCR-06-0226. [DOI] [PubMed] [Google Scholar]
  • 89.Kaelin WG, Jr, Ratcliffe PJ. Oxygen sensing by metazoans: the central role of the HIF hydroxylase pathway. Mol Cell. 2008;30:393–402. doi: 10.1016/j.molcel.2008.04.009. [DOI] [PubMed] [Google Scholar]
  • 90.Kaluz S, Kaluzova M, Stanbridge EJ. Does inhibition of degradation of hypoxia-inducible factor (HIF) alpha always lead to activation of HIF? Lessons learnt from the effect of proteasomal inhibition on HIF activity. J Cell Biochem. 2008;104:536–544. doi: 10.1002/jcb.21644. [DOI] [PubMed] [Google Scholar]
  • 91.Metzen E, Ratcliffe PJ. HIF hydroxylation and cellular oxygen sensing. Biol Chem. 2004;385:223–230. doi: 10.1515/BC.2004.016. [DOI] [PubMed] [Google Scholar]
  • 92.Ivan M, Kondo K, Yang H, Kim W, Valiando J, Ohh M, Salic A, Asara JM, Lane WS, Kaelin WG., Jr HIFalpha targeted for VHL-mediated destruction by proline hydroxylation: implications for O2 sensing. Science. 2001;292:464–468. doi: 10.1126/science.1059817. [DOI] [PubMed] [Google Scholar]
  • 93.Jokilehto T, Jaakkola PM. The role of HIF prolyl hydroxylases in tumour growth. J Cell Mol Med. 2010;14:758–770. doi: 10.1111/j.1582-4934.2010.01030.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Min JH, Yang H, Ivan M, Gertler F, Kaelin WG, Jr, Pavletich NP. Structure of an HIF-1alpha–pVHL complex: hydroxyproline recognition in signaling. Science. 2002;296:1886–1889. doi: 10.1126/science.1073440. [DOI] [PubMed] [Google Scholar]
  • 95.Hewitson KS, McNeill LA, Riordan MV, Tian YM, Bullock AN, Welford RW, Elkins JM, Oldham NJ, Bhattacharya S, Gleadle JM, Ratcliffe PJ, Pugh CW, Schofield CJ. 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;277:26351–26355. doi: 10.1074/jbc.C200273200. [DOI] [PubMed] [Google Scholar]
  • 96.Bertout JA, Patel SA, Simon MC. The impact of O2 availability on human cancer. Nat Rev Cancer. 2008;8:967–975. doi: 10.1038/nrc2540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Chandel NS, Simon MC. Hypoxia-inducible factor: roles in development, physiology, and disease. Cell Death Differ. 2008;15:619–620. doi: 10.1038/cdd.2008.11. [DOI] [PubMed] [Google Scholar]
  • 98.Simon MC, Keith B. The role of oxygen availability in embryonic development and stem cell function. Nat Rev Mol Cell Biol. 2008;9:285–296. doi: 10.1038/nrm2354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Wenger RH. Cellular adaptation to hypoxia: O2-sensing protein hydroxylases, hypoxia-inducible transcription factors, and O2-regulated gene expression. FASEB J. 2002;16:1151–1162. doi: 10.1096/fj.01-0944rev. [DOI] [PubMed] [Google Scholar]
  • 100.Zhong H, De Marzo AM, Laughner E, Lim M, Hilton DA, Zagzag D, Buechler P, Isaacs WB, Semenza GL, Simons JW. Overexpression of hypoxia-inducible factor 1alpha in common human cancers and their metastases. Cancer Res. 1999;59:5830–5835. [PubMed] [Google Scholar]
  • 101.Semenza GL. Regulation of mammalian O2 homeostasis by hypoxia-inducible factor 1. Annu Rev Cell Dev Biol. 1999;15:551–578. doi: 10.1146/annurev.cellbio.15.1.551. [DOI] [PubMed] [Google Scholar]
  • 102.Quintero M, Mackenzie N, Brennan PA. Hypoxia-inducible factor 1 (HIF-1) in cancer. Eur J Surg Oncol. 2004;30:465–468. doi: 10.1016/j.ejso.2004.03.008. [DOI] [PubMed] [Google Scholar]
  • 103.An WG, Kanekal M, Simon MC, Maltepe E, Blagosklonny MV, Neckers LM. Stabilization of wild-type p53 by hypoxia-inducible factor 1alpha. Nature. 1998;392:405–408. doi: 10.1038/32925. [DOI] [PubMed] [Google Scholar]
  • 104.Brugarolas J, Lei K, Hurley RL, Manning BD, Reiling JH, Hafen E, Witters LA, Ellisen LW, Kaelin WG., Jr Regulation of mTOR function in response to hypoxia by REDD1 and the TSC1/TSC2 tumor suppressor complex. Genes Dev. 2004;18:2893–2904. doi: 10.1101/gad.1256804. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Hebert C, Norris K, Parashar P, Ord RA, Nikitakis NG, Sauk JJ. Hypoxia-inducible factor-1alpha polymorphisms and TSC1/2 mutations are complementary in head and neck cancers. Mol Cancer. 2006;5:3. doi: 10.1186/1476-4598-5-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Wolff NC, Vega-Rubin-de-Celis S, Xie XJ, Castrillon DH, Kabbani W, Brugarolas J. Cell-type-dependent regulation of mTORC1 by REDD1 and the tumor suppressors TSC1/TSC2 and LKB1 in response to hypoxia. Mol Cell Biol. 2011;31:1870–1884. doi: 10.1128/MCB.01393-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Sato Y. Persistent vascular normalization as an alternative goal of anti-angiogenic cancer therapy. Cancer Sci. 2011;102:1253–1256. doi: 10.1111/j.1349-7006.2011.01929.x. [DOI] [PubMed] [Google Scholar]
  • 108.Birner P, Schindl M, Obermair A, Plank C, Breitenecker G, Oberhuber G. Overexpression of hypoxia-inducible factor 1alpha is a marker for an unfavorable prognosis in early-stage invasive cervical cancer. Cancer Res. 2000;60:4693–4696. [PubMed] [Google Scholar]
  • 109.Semenza GL. Oxygen sensing, homeostasis, and disease. N Engl J Med. 2011;365:537–547. doi: 10.1056/NEJMra1011165. [DOI] [PubMed] [Google Scholar]
  • 110.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. Br J Cancer. 2007;96:1284–1292. doi: 10.1038/sj.bjc.6603675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Holmquist-Mengelbier L, Fredlund E, Lofstedt T, Noguera R, Navarro S, Nilsson H, Pietras A, Vallon-Christersson J, Borg A, Gradin K, Poellinger L, Pahlman S. 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;10:413–423. doi: 10.1016/j.ccr.2006.08.026. [DOI] [PubMed] [Google Scholar]
  • 112.Wu XH, Qian C, Yuan K. Correlations of hypoxia-inducible factor-1alpha/hypoxia-inducible factor-2alpha expression with angiogenesis factors expression and prognosis in non-small cell lung cancer. Chin Med J (Engl) 2011;124:11–18. [PubMed] [Google Scholar]
  • 113.Kondo K, Kim WY, Lechpammer M, Kaelin WG., Jr Inhibition of HIF2alpha is sufficient to suppress pVHL-defective tumor growth. PLoS Biol. 2003;1:E83. doi: 10.1371/journal.pbio.0000083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Mazumdar J, Hickey MM, Pant DK, Durham AC, Sweet-Cordero A, Vachani A, Jacks T, Chodosh LA, Kissil JL, Simon MC, Keith B. HIF-2alpha deletion promotes Kras-driven lung tumor development. Proc Natl Acad Sci U S A. 2010;107:14182–14187. doi: 10.1073/pnas.1001296107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Helczynska K, Larsson AM, Holmquist ML, Bridges E, Fredlund E, Borgquist S, Landberg G, Pahlman S, Jirstrom K. Hypoxia-inducible factor-2alpha correlates to distant recurrence and poor outcome in invasive breast cancer. Cancer Res. 2008;68:9212–9220. doi: 10.1158/0008-5472.CAN-08-1135. [DOI] [PubMed] [Google Scholar]
  • 116.Kim WY, Perera S, Zhou B, Carretero J, Yeh JJ, Heathcote SA, Jackson AL, Nikolinakos P, Ospina B, Naumov G, Brandstetter KA, Weigman VJ, Zaghlul S, Hayes DN, Padera RF, Heymach JV, Kung AL, Sharpless NE, Kaelin WG, Jr, Wong KK. HIF2alpha cooperates with RAS to promote lung tumorigenesis in mice. J Clin Invest. 2009;119:2160–2170. doi: 10.1172/JCI38443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Li Z, Bao S, Wu Q, Wang H, Eyler C, Sathornsumetee S, Shi Q, Cao Y, Lathia J, McLendon RE, Hjelmeland AB, Rich JN. Hypoxia-inducible factors regulate tumorigenic capacity of glioma stem cells. Cancer Cell. 2009;15:501–513. doi: 10.1016/j.ccr.2009.03.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.McCord AM, Jamal M, Shankavaram UT, Lang FF, Camphausen K, Tofilon PJ. Physiologic oxygen concentration enhances the stem-like properties of CD133+ human glioblastoma cells in vitro. Mol Cancer Res. 2009;7:489–497. doi: 10.1158/1541-7786.MCR-08-0360. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Pietras A, Gisselsson D, Ora I, Noguera R, Beckman S, Navarro S, Pahlman S. High levels of HIF-2alpha highlight an immature neural crest-like neuroblastoma cell cohort located in a perivascular niche. J Pathol. 2008;214:482–488. doi: 10.1002/path.2304. [DOI] [PubMed] [Google Scholar]
  • 120.Pietras A, Johnsson AS, Pahlman S. The HIF-2alpha-driven pseudo-hypoxic phenotype in tumor aggressiveness, differentiation, and vascularization. Curr Top Microbiol Immunol. 2010;345:1–20. doi: 10.1007/82_2010_72. [DOI] [PubMed] [Google Scholar]
  • 121.Rasheed S, Harris AL, Tekkis PP, Turley H, Silver A, McDonald PJ, Talbot IC, Glynne-Jones R, Northover JM, Guenther T. Hypoxia-inducible factor-1alpha and -2alpha are expressed in most rectal cancers but only hypoxia-inducible factor-1alpha is associated with prognosis. Br J Cancer. 2009;100:1666–1673. doi: 10.1038/sj.bjc.6605026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Zhu GQ, Tang YL, Li L, Zheng M, Jiang J, Li XY, Chen SX, Liang XH. Hypoxia inducible factor 1alpha and hypoxia inducible factor 2alpha play distinct and functionally overlapping roles in oral squamous cell carcinoma. Clin Cancer Res. 2010;16:4732–4741. doi: 10.1158/1078-0432.CCR-10-1408. [DOI] [PubMed] [Google Scholar]
  • 123.Osada R, Horiuchi A, Kikuchi N, Yoshida J, Hayashi A, Ota M, Katsuyama Y, Melillo G, Konishi I. Expression of hypoxia-inducible factor 1alpha, hypoxia-inducible factor 2alpha, and von Hippel–Lindau protein in epithelial ovarian neoplasms and allelic loss of von Hippel–Lindau gene: nuclear expression of hypoxia-inducible factor 1alpha is an independent prognostic factor in ovarian carcinoma. Hum Pathol. 2007;38:1310–1320. doi: 10.1016/j.humpath.2007.02.010. [DOI] [PubMed] [Google Scholar]
  • 124.Giatromanolaki A, Sivridis E, Simopoulos C, Polychronidis A, Gatter KC, Harris AL, Koukourakis MI. Hypoxia inducible factors 1alpha and 2alpha are associated with VEGF expression and angiogenesis in gallbladder carcinomas. J Surg Oncol. 2006;94:242–247. doi: 10.1002/jso.20443. [DOI] [PubMed] [Google Scholar]
  • 125.Griffiths EA, Pritchard SA, McGrath SM, Valentine HR, Price PM, Welch IM, West CM. Hypoxia-associated markers in gastric carcinogenesis and HIF-2alpha in gastric and gastro-oesophageal cancer prognosis. Br J Cancer. 2008;98:965–973. doi: 10.1038/sj.bjc.6604210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Winter SC, Shah KA, Han C, Campo L, Turley H, Leek R, Corbridge RJ, Cox GJ, Harris AL. The relation between hypoxia-inducible factor (HIF)-1alpha and HIF-2alpha expression with anemia and outcome in surgically treated head and neck cancer. Cancer. 2006;107:757–766. doi: 10.1002/cncr.21983. [DOI] [PubMed] [Google Scholar]
  • 127.Yoshimura H, Dhar DK, Kohno H, Kubota H, Fujii T, Ueda S, Kinugasa S, Tachibana M, Nagasue N. Prognostic impact of hypoxia-inducible factors 1alpha and 2alpha in colorectal cancer patients: correlation with tumor angiogenesis and cyclooxygenase-2 expression. Clin Cancer Res. 2004;10:8554–8560. doi: 10.1158/1078-0432.CCR-0946-03. [DOI] [PubMed] [Google Scholar]
  • 128.Heidbreder M, Frohlich F, Johren O, Dendorfer A, Qadri F, Dominiak P. Hypoxia rapidly activates HIF-3alpha mRNA expression. FASEB J. 2003;17:1541–1543. doi: 10.1096/fj.02-0963fje. [DOI] [PubMed] [Google Scholar]
  • 129.Heidbreder M, Qadri F, Johren O, Dendorfer A, Depping R, Frohlich F, Wagner KF, Dominiak P. Non-hypoxic induction of HIF-3alpha by 2-deoxy-d-glucose and insulin. Biochem Biophys Res Commun. 2007;352:437–443. doi: 10.1016/j.bbrc.2006.11.027. [DOI] [PubMed] [Google Scholar]
  • 130.Makino Y, Kanopka A, Wilson WJ, Tanaka H, Poellinger L. Inhibitory PAS domain protein (IPAS) is a hypoxia-inducible splicing variant of the hypoxia-inducible factor-3alpha locus. J Biol Chem. 2002;277:32405–32408. doi: 10.1074/jbc.C200328200. [DOI] [PubMed] [Google Scholar]
  • 131.Maynard MA, Qi H, Chung J, Lee EH, Kondo Y, Hara S, Conaway RC, Conaway JW, Ohh M. Multiple splice variants of the human HIF-3 alpha locus are targets of the von Hippel–Lindau E3 ubiquitin ligase complex. J Biol Chem. 2003;278:11032–11040. doi: 10.1074/jbc.M208681200. [DOI] [PubMed] [Google Scholar]
  • 132.Augstein A, Poitz DM, Braun-Dullaeus RC, Strasser RH, Schmeisser A. Cell-specific and hypoxia-dependent regulation of human HIF-3alpha: inhibition of the expression of HIF target genes in vascular cells. Cell Mol Life Sci. 2011;68:2627–2642. doi: 10.1007/s00018-010-0575-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Maynard MA, Evans AJ, Shi W, Kim WY, Liu FF, Ohh M. Dominant-negative HIF-3 alpha 4 suppresses VHL-null renal cell carcinoma progression. Cell Cycle. 2007;6:2810–2816. doi: 10.4161/cc.6.22.4947. [DOI] [PubMed] [Google Scholar]
  • 134.Hermann PC, Bhaskar S, Cioffi M, Heeschen C. Cancer stem cells in solid tumors. Semin Cancer Biol. 2010;20:77–84. doi: 10.1016/j.semcancer.2010.03.004. [DOI] [PubMed] [Google Scholar]
  • 135.Ischenko I, Seeliger H, Kleespies A, Angele MK, Eichhorn ME, Jauch KW, Bruns CJ. Pancreatic cancer stem cells: new understanding of tumorigenesis, clinical implications. Langenbecks Arch Surg. 2010;395:1–10. doi: 10.1007/s00423-009-0502-z. [DOI] [PubMed] [Google Scholar]
  • 136.Lee CJ, Dosch J, Simeone DM. Pancreatic cancer stem cells. J Clin Oncol. 2008;26:2806–2812. doi: 10.1200/JCO.2008.16.6702. [DOI] [PubMed] [Google Scholar]
  • 137.Sarkar FH, Li Y, Wang Z, Kong D. Pancreatic cancer stem cells and EMT in drug resistance and metastasis. Minerva Chir. 2009;64:489–500. [PMC free article] [PubMed] [Google Scholar]
  • 138.Creighton CJ, Chang JC, Rosen JM. Epithelial–mesenchymal transition (EMT) in tumor-initiating cells and its clinical implications in breast cancer. J Mammary Gland Biol Neoplasia. 2010;15:253–260. doi: 10.1007/s10911-010-9173-1. [DOI] [PubMed] [Google Scholar]
  • 139.Wang Z, Li Y, Ahmad A, Azmi AS, Kong D, Banerjee S, Sarkar FH. Targeting miRNAs involved in cancer stem cell and EMT regulation: an emerging concept in overcoming drug resistance. Drug Resist Updat. 2010;13:109–118. doi: 10.1016/j.drup.2010.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Wang Z, Li Y, Ahmad A, Banerjee S, Azmi AS, Kong D, Sarkar FH. Pancreatic cancer: understanding and overcoming chemoresistance. Nat Rev Gastroenterol Hepatol. 2011;8:27–33. doi: 10.1038/nrgastro.2010.188. [DOI] [PubMed] [Google Scholar]
  • 141.Fraker CA, Ricordi C, Inverardi L, Dominguez-Bendala J. Oxygen: a master regulator of pancreatic development? Biol Cell. 2009;101:431–440. doi: 10.1042/BC20080178. [DOI] [PubMed] [Google Scholar]
  • 142.Dunwoodie SL. The role of hypoxia in development of the mammalian embryo. Dev Cell. 2009;17:755–773. doi: 10.1016/j.devcel.2009.11.008. [DOI] [PubMed] [Google Scholar]
  • 143.Ezashi T, Das P, Roberts RM. Low O2 tensions and the prevention of differentiation of hES cells. Proc Natl Acad Sci U S A. 2005;102:4783–4788. doi: 10.1073/pnas.0501283102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Gustafsson MV, Zheng X, Pereira T, Gradin K, Jin S, Lundkvist J, Ruas JL, Poellinger L, Lendahl U, Bondesson M. Hypoxia requires notch signaling to maintain the undifferentiated cell state. Dev Cell. 2005;9:617–628. doi: 10.1016/j.devcel.2005.09.010. [DOI] [PubMed] [Google Scholar]
  • 145.Chen HL, Pistollato F, Hoeppner DJ, Ni HT, McKay RD, Panchision DM. Oxygen tension regulates survival and fate of mouse central nervous system precursors at multiple levels. Stem Cells. 2007;25:2291–2301. doi: 10.1634/stemcells.2006-0609. [DOI] [PubMed] [Google Scholar]
  • 146.Soeda A, Park M, Lee D, Mintz A, ndroutsellis-Theotokis A, McKay RD, Engh J, Iwama T, Kunisada T, Kassam AB, Pollack IF, Park DM. Hypoxia promotes expansion of the CD133-positive glioma stem cells through activation of HIF-1alpha. Oncogene. 2009;28:3949–3959. doi: 10.1038/onc.2009.252. [DOI] [PubMed] [Google Scholar]
  • 147.Panchision DM. The role of oxygen in regulating neural stem cells in development and disease. J Cell Physiol. 2009;220:562–568. doi: 10.1002/jcp.21812. [DOI] [PubMed] [Google Scholar]
  • 148.Zeng W, Wan R, Zheng Y, Singh SR, Wei Y. Hypoxia, stem cells and bone tumor. Cancer Lett. 2011;313:129–136. doi: 10.1016/j.canlet.2011.09.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Simsek T, Kocabas F, Zheng J, DeBerardinis RJ, Mahmoud AI, Olson EN, Schneider JW, Zhang CC, Sadek HA. The distinct metabolic profile of hematopoietic stem cells reflects their location in a hypoxic niche. Cell Stem Cell. 2010;7:380–390. doi: 10.1016/j.stem.2010.07.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Heddleston JM, Li Z, Lathia JD, Bao S, Hjelmeland AB, Rich JN. Hypoxia inducible factors in cancer stem cells. Br J Cancer. 2010;102:789–795. doi: 10.1038/sj.bjc.6605551. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Li Z, Rich JN. Hypoxia and hypoxia inducible factors in cancer stem cell maintenance. Curr Top Microbiol Immunol. 2010;345:21–30. doi: 10.1007/82_2010_75. [DOI] [PubMed] [Google Scholar]
  • 152.Mathieu J, Zhang Z, Zhou W, Wang AJ, Heddleston JM, Pinna CM, Hubaud A, Stadler B, Choi M, Bar M, Tewari M, Liu A, Vessella R, Rostomily R, Born D, Horwitz M, Ware C, Blau CA, Cleary MA, Rich JN, Ruohola-Baker H. HIF induces human embryonic stem cell markers in cancer cells. Cancer Res. 2011;71:4640–4652. doi: 10.1158/0008-5472.CAN-10-3320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Covello KL, Kehler J, Yu H, Gordan JD, Arsham AM, Hu CJ, Labosky PA, Simon MC, Keith B. HIF-2alpha regulates Oct-4: effects of hypoxia on stem cell function, embryonic development, and tumor growth. Genes Dev. 2006;20:557–570. doi: 10.1101/gad.1399906. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Gordan JD, Lal P, Dondeti VR, Letrero R, Parekh KN, Oquendo CE, Greenberg RA, Flaherty KT, Rathmell WK, Keith B, Simon MC, Nathanson KL. HIF-alpha effects on c-Myc distinguish two subtypes of sporadic VHL-deficient clear cell renal carcinoma. Cancer Cell. 2008;14:435–446. doi: 10.1016/j.ccr.2008.10.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Li L, Bhatia R. Stem cell quiescence. Clin Cancer Res. 2011;17:4936–4941. doi: 10.1158/1078-0432.CCR-10-1499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Wang Y, Liu Y, Malek SN, Zheng P, Liu Y. Targeting HIF1alpha eliminates cancer stem cells in hematological malignancies. Cell Stem Cell. 2011;8:399–411. doi: 10.1016/j.stem.2011.02.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Peurala E, Koivunen P, Bloigu R, Haapasaari KM, Jukkola-Vuorinen A. Expressions of individual PHDs associate with good prognostic factors and increased proliferation in breast cancer patients. Breast Cancer Res Treat. 2012;133:179–188. doi: 10.1007/s10549-011-1750-5. [DOI] [PubMed] [Google Scholar]
  • 158.Bar EE. Glioblastoma, cancer stem cells and hypoxia. Brain Pathol. 2011;21:119–129. doi: 10.1111/j.1750-3639.2010.00460.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Heddleston JM, Li Z, McLendon RE, Hjelmeland AB, Rich JN. The hypoxic microenvironment maintains glioblastoma stem cells and promotes reprogramming towards a cancer stem cell phenotype. Cell Cycle. 2009;8:3274–3284. doi: 10.4161/cc.8.20.9701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Skuli N, Liu L, Runge A, Wang T, Yuan L, Patel S, Iruela-Arispe L, Simon MC, Keith B. Endothelial deletion of hypoxia-inducible factor-2alpha (HIF-2alpha) alters vascular function and tumor angiogenesis. Blood. 2009;114:469–477. doi: 10.1182/blood-2008-12-193581. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Hjelmeland AB, Wu Q, Heddleston JM, Choudhary GS, MacSwords J, Lathia JD, McLendon R, Lindner D, Sloan A, Rich JN. Acidic stress promotes a glioma stem cell phenotype. Cell Death Differ. 2011;18:829–840. doi: 10.1038/cdd.2010.150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Bar EE, Lin A, Mahairaki V, Matsui W, Eberhart CG. Hypoxia increases the expression of stem-cell markers and promotes clonogenicity in glioblastoma neurospheres. Am J Pathol. 2010;177:1491–1502. doi: 10.2353/ajpath.2010.091021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Holmquist L, Lofstedt T, Pahlman S. Effect of hypoxia on the tumor phenotype: the neuroblastoma and breast cancer models. Adv Exp Med Biol. 2006;587:179–193. doi: 10.1007/978-1-4020-5133-3_16. [DOI] [PubMed] [Google Scholar]
  • 164.Koh MY, Darnay BG, Powis G. Hypoxia-associated factor, a novel E3-ubiquitin ligase, binds and ubiquitinates hypoxia-inducible factor 1alpha, leading to its oxygen-independent degradation. Mol Cell Biol. 2008;28:7081–7095. doi: 10.1128/MCB.00773-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Koh MY, Lemos R, Jr, Liu X, Powis G. The hypoxia-associated factor switches cells from HIF-1alpha- to HIF-2alpha-dependent signaling promoting stem cell characteristics, aggressive tumor growth and invasion. Cancer Res. 2011;71:4015–4027. doi: 10.1158/0008-5472.CAN-10-4142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Wang Z, Li Y, Kong D, Sarkar FH. The role of Notch signaling pathway in epithelial–mesenchymal transition (EMT) during development and tumor aggressiveness. Curr Drug Targets. 2010;11:745–751. doi: 10.2174/138945010791170860. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Hugo H, Ackland ML, Blick T, Lawrence MG, Clements JA, Williams ED, Thompson EW. Epithelial–mesenchymal and mesenchymal–epithelial transitions in carcinoma progression. J Cell Physiol. 2007;213:374–383. doi: 10.1002/jcp.21223. [DOI] [PubMed] [Google Scholar]
  • 168.Haase VH. Oxygen regulates epithelial-to-mesenchymal transition: insights into molecular mechanisms and relevance to disease. Kidney Int. 2009;76:492–499. doi: 10.1038/ki.2009.222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Lee JM, Dedhar S, Kalluri R, Thompson EW. The epithelial–mesenchymal transition: new insights in signaling, development, and disease. J Cell Biol. 2006;172:973–981. doi: 10.1083/jcb.200601018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Jiang J, Tang YL, Liang XH. EMT: a new vision of hypoxia promoting cancer progression. Cancer Biol Ther. 2011;11:714–723. doi: 10.4161/cbt.11.8.15274. [DOI] [PubMed] [Google Scholar]
  • 171.Graham TR, Zhau HE, Odero-Marah VA, Osunkoya AO, Kimbro KS, Tighiouart M, Liu T, Simons JW, O’Regan RM. Insulin-like growth factor-Idependent up-regulation of ZEB1 drives epithelial-to-mesenchymal transition in human prostate cancer cells. Cancer Res. 2008;68:2479–2488. doi: 10.1158/0008-5472.CAN-07-2559. [DOI] [PubMed] [Google Scholar]
  • 172.Christiansen JJ, Rajasekaran AK. Reassessing epithelial to mesenchymal transition as a prerequisite for carcinoma invasion and metastasis. Cancer Res. 2006;66:8319–8326. doi: 10.1158/0008-5472.CAN-06-0410. [DOI] [PubMed] [Google Scholar]
  • 173.Eger A, Stockinger A, Schaffhauser B, Beug H, Foisner R. Epithelial mesenchymal transition by c-Fos estrogen receptor activation involves nuclear translocation of beta-catenin and upregulation of beta-catenin/lymphoid enhancer binding factor-1 transcriptional activity. J Cell Biol. 2000;148:173–188. doi: 10.1083/jcb.148.1.173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Jing Y, Han Z, Zhang S, Liu Y, Wei L. Epithelial–mesenchymal transition in tumor microenvironment. Cell Biosci. 2011;1:29. doi: 10.1186/2045-3701-1-29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Cannito S, Novo E, di Bonzo LV, Busletta C, Colombatto S, Parola M. Epithelial– mesenchymal transition: from molecular mechanisms, redox regulation to implications in human health and disease. Antioxid Redox Signal. 2010;12:1383–1430. doi: 10.1089/ars.2009.2737. [DOI] [PubMed] [Google Scholar]
  • 176.Chang Q, Jurisica I, Do T, Hedley DW. Hypoxia predicts aggressive growth and spontaneous metastasis formation from orthotopically grown primary xenografts of human pancreatic cancer. Cancer Res. 2011;71:3110–3120. doi: 10.1158/0008-5472.CAN-10-4049. [DOI] [PubMed] [Google Scholar]
  • 177.Luo Y, He DL, Ning L, Shen SL, Li L, Li X. Hypoxia-inducible factor-1alpha induces the epithelial–mesenchymal transition of human prostatecancer cells. Chin Med J (Engl) 2006;119:713–718. [PubMed] [Google Scholar]
  • 178.Sahlgren C, Gustafsson MV, Jin S, Poellinger L, Lendahl U. Notch signaling mediates hypoxia-induced tumor cell migration and invasion. Proc Natl Acad Sci U S A. 2008;105:6392–6397. doi: 10.1073/pnas.0802047105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Yang MH, Wu MZ, Chiou SH, Chen PM, Chang SY, Liu CJ, Teng SC, Wu KJ. Direct regulation of TWIST by HIF-1alpha promotes metastasis. Nat Cell Biol. 2008;10:295–305. doi: 10.1038/ncb1691. [DOI] [PubMed] [Google Scholar]
  • 180.Klymkowsky MW, Savagner P. Epithelial–mesenchymal transition: a cancer researcher’s conceptual friend and foe. Am J Pathol. 2009;174:1588–1593. doi: 10.2353/ajpath.2009.080545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Sun S, Ning X, Zhang Y, Lu Y, Nie Y, Han S, Liu L, Du R, Xia L, He L, Fan D. Hypoxia-inducible factor-1alpha induces Twist expression in tubular epithelial cells subjected to hypoxia, leading to epithelial-to-mesenchymal transition. Kidney Int. 2009;75:1278–1287. doi: 10.1038/ki.2009.62. [DOI] [PubMed] [Google Scholar]
  • 182.Zhou G, Dada LA, Wu M, Kelly A, Trejo H, Zhou Q, Varga J, Sznajder JI. Hypoxia-induced alveolar epithelial–mesenchymal transition requires mitochondrial ROS and hypoxia-inducible factor 1. Am J Physiol Lung Cell Mol Physiol. 2009;297:L1120–L1130. doi: 10.1152/ajplung.00007.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Welford SM, Giaccia AJ. Hypoxia and senescence: the impact of oxygenation on tumor suppression. Mol Cancer Res. 2011;9:538–544. doi: 10.1158/1541-7786.MCR-11-0065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Katoh M, Katoh M. Integrative genomic analyses of ZEB2: transcriptional regulation of ZEB2 based on SMADs, ETS1, HIF1alpha, POU/OCT, and NF-kappaB. Int J Oncol. 2009;34:1737–1742. doi: 10.3892/ijo_00000304. [DOI] [PubMed] [Google Scholar]
  • 185.Dashwood WM, Orner GA, Dashwood RH. Inhibition of beta-catenin/Tcf activity by white tea, green tea, and epigallocatechin-3-gallate (EGCG): minor contribution of H2O2 at physiologically relevant EGCG concentrations. Biochem Biophys Res Commun. 2002;296:584–588. doi: 10.1016/s0006-291x(02)00914-2. [DOI] [PubMed] [Google Scholar]
  • 186.Fan JM, Ng YY, Hill PA, Nikolic-Paterson DJ, Mu W, Atkins RC, Lan HY. Transforming growth factor-beta regulates tubular epithelial–myofibroblast transdifferentiation in vitro. Kidney Int. 1999;56:1455–1467. doi: 10.1046/j.1523-1755.1999.00656.x. [DOI] [PubMed] [Google Scholar]
  • 187.Gal A, Sjoblom T, Fedorova L, Imreh S, Beug H, Moustakas A. Sustained TGF beta exposure suppresses Smad and non-Smad signalling in mammary epithelial cells, leading to EMT and inhibition of growth arrest and apoptosis. Oncogene. 2008;27:1218–1230. doi: 10.1038/sj.onc.1210741. [DOI] [PubMed] [Google Scholar]
  • 188.Hales AM, Schulz MW, Chamberlain CG, McAvoy JW. TGF-beta 1 induces lens cells to accumulate alpha-smooth muscle actin, a marker for subcapsular cataracts. Curr Eye Res. 1994;13:885–890. doi: 10.3109/02713689409015091. [DOI] [PubMed] [Google Scholar]
  • 189.Higgins DF, Biju MP, Akai Y, Wutz A, Johnson RS, Haase VH. Hypoxic induction of Ctgf is directly mediated by Hif-1. Am J Physiol Renal Physiol. 2004;287:F1223–F1232. doi: 10.1152/ajprenal.00245.2004. [DOI] [PubMed] [Google Scholar]
  • 190.Kadesch T. Notch signaling: the demise of elegant simplicity. Curr Opin Genet Dev. 2004;14:506–512. doi: 10.1016/j.gde.2004.07.007. [DOI] [PubMed] [Google Scholar]
  • 191.Keeton MR, Curriden SA, van Zonneveld AJ, Loskutoff DJ. Identification of regulatory sequences in the type 1 plasminogen activator inhibitor gene responsive to transforming growth factor beta. J Biol Chem. 1991;266:23048–23052. [PubMed] [Google Scholar]
  • 192.Kietzmann T, Roth U, Jungermann K. Induction of the plasminogen activator inhibitor-1 gene expression by mild hypoxia via a hypoxia response element binding the hypoxia-inducible factor-1 in rat hepatocytes. Blood. 1999;94:4177–4185. [PubMed] [Google Scholar]
  • 193.Kim K, Lu Z, Hay ED. Direct evidence for a role of beta-catenin/LEF-1 signaling pathway in induction of EMT. Cell Biol Int. 2002;26:463–476. doi: 10.1006/cbir.2002.0901. [DOI] [PubMed] [Google Scholar]
  • 194.Lee JJ, Natsuizaka M, Ohashi S, Wong GS, Takaoka M, Michaylira CZ, Budo D, Tobias JW, Kanai M, Shirakawa Y, Naomoto Y, Klein-Szanto AJ, Haase VH, Nakagawa H. Hypoxia activates the cyclooxygenase-2-prostaglandin E synthase axis. Carcinogenesis. 2010;31:427–434. doi: 10.1093/carcin/bgp326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Lee YH, Albig AR, Regner M, Schiemann BJ, Schiemann WP. Fibulin-5 initiates epithelial–mesenchymal transition (EMT) and enhances EMT induced by TGF-beta in mammary epithelial cells via a MMP-dependent mechanism. Carcinogenesis. 2008;29:2243–2251. doi: 10.1093/carcin/bgn199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Leivonen SK, Kahari VM. Transforming growth factor-beta signaling in cancer invasion and metastasis. Int J Cancer. 2007;121:2119–2124. doi: 10.1002/ijc.23113. [DOI] [PubMed] [Google Scholar]
  • 197.Li Y, Wang Z, Kong D, Li R, Sarkar SH, Sarkar FH. Regulation of Akt/FOXO3a/GSK-3beta/AR signaling network by isoflavone in prostate cancer cells. J Biol Chem. 2008;283:27707–27716. doi: 10.1074/jbc.M802759200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.McMahon S, Charbonneau M, Grandmont S, Richard DE, Dubois CM. Transforming growth factor beta1 induces hypoxia-inducible factor-1 stabilization through selective inhibition of PHD2 expression. J Biol Chem. 2006;281:24171–24181. doi: 10.1074/jbc.M604507200. [DOI] [PubMed] [Google Scholar]
  • 199.Miettinen PJ, Ebner R, Lopez AR, Derynck R. TGF-beta induced transdifferentiation of mammary epithelial cells to mesenchymal cells: involvement of type I receptors. J Cell Biol. 1994;127:2021–2036. doi: 10.1083/jcb.127.6.2021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Neil JR, Johnson KM, Nemenoff RA, Schiemann WP. Cox-2 inactivates Smad signaling and enhances EMT stimulated by TGF-beta through a PGE2-dependent mechanisms. Carcinogenesis. 2008;29:2227–2235. doi: 10.1093/carcin/bgn202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Nishi H, Nakada T, Hokamura M, Osakabe Y, Itokazu O, Huang LE, Isaka K. Hypoxia-inducible factor-1 transactivates transforming growth factor-beta3 in trophoblast. Endocrinology. 2004;145:4113–4118. doi: 10.1210/en.2003-1639. [DOI] [PubMed] [Google Scholar]
  • 202.Saad S, Stanners SR, Yong R, Tang O, Pollock CA. Notch mediated epithelial to mesenchymal transformation is associated with increased expression of the Snail transcription factor. Int J Biochem Cell Biol. 2010;42:1115–1122. doi: 10.1016/j.biocel.2010.03.016. [DOI] [PubMed] [Google Scholar]
  • 203.Saika S, Kono-Saika S, Ohnishi Y, Sato M, Muragaki Y, Ooshima A, Flanders KC, Yoo J, Anzano M, Liu CY, Kao WW, Roberts AB. Smad3 signaling is required for epithelial–mesenchymal transition of lens epithelium after injury. Am J Pathol. 2004;164:651–663. doi: 10.1016/S0002-9440(10)63153-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Schaffer L, Scheid A, Spielmann P, Breymann C, Zimmermann R, Meuli M, Gassmann M, Marti HH, Wenger RH. Oxygen-regulated expression of TGF-beta 3, a growth factor involved in trophoblast differentiation. Placenta. 2003;24:941–950. doi: 10.1016/s0143-4004(03)00166-8. [DOI] [PubMed] [Google Scholar]
  • 205.Taylor MA, Parvani JG, Schiemann WP. The pathophysiology of epithelial–mesenchymal transition induced by transforming growth factor-beta in normal and malignant mammary epithelial cells. J Mammary Gland Biol Neoplasia. 2010;15:169–190. doi: 10.1007/s10911-010-9181-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Vincent T, Neve EP, Johnson JR, Kukalev A, Rojo F, Albanell J, Pietras K, Virtanen I, Philipson L, Leopold PL, Crystal RG, de Herreros AG, Moustakas A, Pettersson RF, Fuxe J. A SNAIL1–SMAD3/4 transcriptional repressor complex promotes TGF-beta mediated epithelial–mesenchymal transition. Nat Cell Biol. 2009;11:943–950. doi: 10.1038/ncb1905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Wagner J, Lehmann L. Estrogens modulate the gene expression of Wnt-7a in cultured endometrial adenocarcinoma cells. Mol Nutr Food Res. 2006;50:368–372. doi: 10.1002/mnfr.200500215. [DOI] [PubMed] [Google Scholar]
  • 208.Willis BC, Liebler JM, Luby-Phelps K, Nicholson AG, Crandall ED, du Bois RM, Borok Z. Induction of epithelial–mesenchymal transition in alveolar epithelial cells by transforming growth factor-beta1: potential role in idiopathic pulmonary fibrosis. Am J Pathol. 2005;166:1321–1332. doi: 10.1016/s0002-9440(10)62351-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Yang L, Lin C, Liu ZR. P68 RNA helicase mediates PDGF-induced epithelial mesenchymal transition by displacing Axin from beta-catenin. Cell. 2006;127:139–155. doi: 10.1016/j.cell.2006.08.036. [DOI] [PubMed] [Google Scholar]
  • 210.Zavadil J, Cermak L, Soto-Nieves N, Bottinger EP. Integration of TGF-beta/Smad and Jagged1/Notch signalling in epithelial-to-mesenchymal transition. EMBO J. 2004;23:1155–1165. doi: 10.1038/sj.emboj.7600069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Zavadil J, Bottinger EP. TGF-beta and epithelial-to-mesenchymal transitions. Oncogene. 2005;24:5764–5774. doi: 10.1038/sj.onc.1208927. [DOI] [PubMed] [Google Scholar]
  • 212.Zhang C, Meng X, Zhu Z, Liu J, Deng A. Connective tissue growth factor regulates the key events in tubular epithelial to myofibroblast transition in vitro. Cell Biol Int. 2004;28:863–873. doi: 10.1016/j.cellbi.2004.09.003. [DOI] [PubMed] [Google Scholar]
  • 213.Huber MA, Azoitei N, Baumann B, Grunert S, Sommer A, Pehamberger H, Kraut N, Beug H, Wirth T. NF-kappaB is essential for epithelial–mesenchymal transition and metastasis in a model of breast cancer progression. J Clin Invest. 2004;114:569–581. doi: 10.1172/JCI21358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Chuang MJ, Sun KH, Tang SJ, Deng MW, Wu YH, Sung JS, Cha TL, Sun GH. Tumor-derived tumor necrosis factor-alpha promotes progression and epithelial– mesenchymal transition in renal cell carcinoma cells. Cancer Sci. 2008;99:905–913. doi: 10.1111/j.1349-7006.2008.00756.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Sullivan NJ, Sasser AK, Axel AE, Vesuna F, Raman V, Ramirez N, Oberyszyn TM, Hall BM. Interleukin-6 induces an epithelial–mesenchymal transition phenotype in human breast cancer cells. Oncogene. 2009;28:2940–2947. doi: 10.1038/onc.2009.180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.St John MA, Dohadwala M, Luo J, Wang G, Lee G, Shih H, Heinrich E, Krysan K, Walser T, Hazra S, Zhu L, Lai C, Abemayor E, Fishbein M, Elashoff DA, Sharma S, Dubinett SM. Proinflammatory mediators upregulate snail in head and neck squamous cell carcinoma. Clin Cancer Res. 2009;15:6018–6027. doi: 10.1158/1078-0432.CCR-09-0011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Chen J, Imanaka N, Chen J, Griffin JD. Hypoxia potentiates Notch signaling in breast cancer leading to decreased E-cadherin expression and increased cell migration and invasion. Br J Cancer. 2010;102:351–360. doi: 10.1038/sj.bjc.6605486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Erler JT, Bennewith KL, Nicolau M, Dornhofer N, Kong C, Le QT, Chi JT, Jeffrey SS, Giaccia AJ. Lysyl oxidase is essential for hypoxia-induced metastasis. Nature. 2006;440:1222–1226. doi: 10.1038/nature04695. [DOI] [PubMed] [Google Scholar]
  • 219.Gupta R, Chetty C, Bhoopathi P, Lakka S, Mohanam S, Rao JS, Dinh DE. Downregulation of uPA/uPAR inhibits intermittent hypoxia-induced epithelial– mesenchymal transition (EMT) in DAOY and D283 medulloblastoma cells. Int J Oncol. 2011;38:733–744. doi: 10.3892/ijo.2010.883. [DOI] [PubMed] [Google Scholar]
  • 220.Gonzalez-Moreno O, Lecanda J, Green JE, Segura V, Catena R, Serrano D, Calvo A. VEGF elicits epithelial–mesenchymal transition (EMT) in prostate intra-epithelial neoplasia (PIN)-like cells via an autocrine loop. Exp Cell Res. 2010;316:554–567. doi: 10.1016/j.yexcr.2009.11.020. [DOI] [PubMed] [Google Scholar]
  • 221.Nesbitt TL, Roberts A, Tan H, Junor L, Yost MJ, Potts JD, Dettman RW, Goodwin RL. Coronary endothelial proliferation and morphogenesis are regulated by a VEGF-mediated pathway. Dev Dyn. 2009;238:423–430. doi: 10.1002/dvdy.21847. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Paruchuri S, Yang JH, Aikawa E, Melero-Martin JM, Khan ZA, Loukogeorgakis S, Schoen FJ, Bischoff J. Human pulmonary valve progenitor cells exhibit endothelial/mesenchymal plasticity in response to vascular endothelial growth factor-A and transforming growth factor-beta2. Circ Res. 2006;99:861–869. doi: 10.1161/01.RES.0000245188.41002.2c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Yang AD, Camp ER, Fan F, Shen L, Gray MJ, Liu W, Somcio R, Bauer TW, Wu Y, Hicklin DJ, Ellis LM. Vascular endothelial growth factor receptor-1 activation mediates epithelial to mesenchymal transition in human pancreatic carcinoma cells. Cancer Res. 2006;66:46–51. doi: 10.1158/0008-5472.CAN-05-3086. [DOI] [PubMed] [Google Scholar]
  • 224.Wu MZ, Tsai YP, Yang MH, Huang CH, Chang SY, Chang CC, Teng SC, Wu KJ. Interplay between HDAC3 and WDR5 is essential for hypoxia-induced epithelial–mesenchymal transition. Mol Cell. 2011;43:811–822. doi: 10.1016/j.molcel.2011.07.012. [DOI] [PubMed] [Google Scholar]
  • 225.An FQ, Matsuda M, Fujii H, Matsumoto Y. Expression of vascular endothelial growth factor in surgical specimens of hepatocellular carcinoma. J Cancer Res Clin Oncol. 2000;126:153–160. doi: 10.1007/s004320050025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Yi ZY, Feng LJ, Xiang Z, Yao H. Vascular endothelial growth factor receptor-1 activation mediates epithelial to mesenchymal transition in hepatocellular carcinoma cells. J Invest Surg. 2011;24:67–76. doi: 10.3109/08941939.2010.542272. [DOI] [PubMed] [Google Scholar]
  • 227.Li Y, Yang X, Su LJ, Flaig TW. VEGFR and EGFR inhibition increases epithelial cellular characteristics and chemotherapy sensitivity in mesenchymal bladder cancer cells. Oncol Rep. 2010;24:1019–1028. doi: 10.3892/or.2010.1019. [DOI] [PubMed] [Google Scholar]
  • 228.Tang J, Wang J, Kong X, Yang J, Guo L, Zheng F, Zhang L, Huang Y, Wan Y. Vascular endothelial growth factor promotes cardiac stem cell migration via the PI3K/Akt pathway. Exp Cell Res. 2009;315:3521–3531. doi: 10.1016/j.yexcr.2009.09.026. [DOI] [PubMed] [Google Scholar]
  • 229.Bao S, Wu Q, Sathornsumetee S, Hao Y, Li Z, Hjelmeland AB, Shi Q, McLendon RE, Bigner DD, Rich JN. Stem cell-like glioma cells promote tumor angiogenesis through vascular endothelial growth factor. Cancer Res. 2006;66:7843–7848. doi: 10.1158/0008-5472.CAN-06-1010. [DOI] [PubMed] [Google Scholar]
  • 230.Cummins EP, Taylor CT. Hypoxia-responsive transcription factors. Pflugers Arch. 2005;450:363–371. doi: 10.1007/s00424-005-1413-7. [DOI] [PubMed] [Google Scholar]
  • 231.Koong AC, Chen EY, Giaccia AJ. Hypoxia causes the activation of nuclear factor kappa B through the phosphorylation of I kappa B alpha on tyrosine residues. Cancer Res. 1994;54:1425–1430. [PubMed] [Google Scholar]
  • 232.Barnes PJ. Nuclear factor-kappa B. Int J Biochem Cell Biol. 1997;29:867–870. doi: 10.1016/s1357-2725(96)00159-8. [DOI] [PubMed] [Google Scholar]
  • 233.Perkins ND. Achieving transcriptional specificity with NF-kappa B. Int J Biochem Cell Biol. 1997;29:1433–1448. doi: 10.1016/s1357-2725(97)00088-5. [DOI] [PubMed] [Google Scholar]
  • 234.Arlt A, Gehrz A, Muerkoster S, Vorndamm J, Kruse ML, Folsch UR, Schafer H. Role of NF-kappaB and Akt/PI3K in the resistance of pancreatic carcinoma cell lines against gemcitabine-induced cell death. Oncogene. 2003;22:3243–3251. doi: 10.1038/sj.onc.1206390. [DOI] [PubMed] [Google Scholar]
  • 235.Sebens S, Arlt A, Schafer H. NF-kappaB as a molecular target in the therapy of pancreatic carcinoma. Recent Results Cancer Res. 2008;177:151–164. doi: 10.1007/978-3-540-71279-4_17. [DOI] [PubMed] [Google Scholar]
  • 236.Zhang Z, Rigas B. NF-kappaB, inflammation and pancreatic carcinogenesis: NF-kappaB as a chemoprevention target (review) Int J Oncol. 2006;29:185–192. [PubMed] [Google Scholar]
  • 237.Alison MR, Nicholson LJ, Lin WR. Chronic inflammation and hepatocellular carcinoma. Recent Results Cancer Res. 2011;185:135–148. doi: 10.1007/978-3-642-03503-6_8. [DOI] [PubMed] [Google Scholar]
  • 238.Wang S, Liu Z, Wang L, Zhang X. NF-kappaB signaling pathway, inflammation and colorectal cancer. Cell Mol Immunol. 2009;6:327–334. doi: 10.1038/cmi.2009.43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239.Sarkar FH, Li Y, Wang Z, Kong D. NF-kappaB signaling pathway and its therapeutic implications in human diseases. Int Rev Immunol. 2008;27:293–319. doi: 10.1080/08830180802276179. [DOI] [PubMed] [Google Scholar]
  • 240.Widera D, Kaus A, Kaltschmidt C, Kaltschmidt B. Neural stem cells, inflammation and NF-kappaB: basic principle of maintenance and repair or origin of brain tumours? J Cell Mol Med. 2008;12:459–470. doi: 10.1111/j.1582-4934.2007.00208.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Taylor CT. Interdependent roles for hypoxia inducible factor and nuclear factor-kappaB in hypoxic inflammation. J Physiol. 2008;586:4055–4059. doi: 10.1113/jphysiol.2008.157669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Taylor CT, Cummins EP. The role of NF-kappaB in hypoxia-induced gene expression. Ann N Y Acad Sci. 2009;1177:178–184. doi: 10.1111/j.1749-6632.2009.05024.x. [DOI] [PubMed] [Google Scholar]
  • 243.Walmsley SR, Print C, Farahi N, Peyssonnaux C, Johnson RS, Cramer T, Sobolewski A, Condliffe AM, Cowburn AS, Johnson N, Chilvers ER. Hypoxia-induced neutrophil survival is mediated by HIF-1alpha-dependent NF-kappaB activity. J Exp Med. 2005;201:105–115. doi: 10.1084/jem.20040624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Scortegagna M, Cataisson C, Martin RJ, Hicklin DJ, Schreiber RD, Yuspa SH, Arbeit JM. HIF-1alpha regulates epithelial inflammation by cell autonomous NFkappaB activation and paracrine stromal remodeling. Blood. 2008;111:3343–3354. doi: 10.1182/blood-2007-10-115758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Culver C, Sundqvist A, Mudie S, Melvin A, Xirodimas D, Rocha S. Mechanism of hypoxia-induced NF-kappaB. Mol Cell Biol. 2010;30:4901–4921. doi: 10.1128/MCB.00409-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Nam SY, Ko YS, Jung J, Yoon J, Kim YH, Choi YJ, Park JW, Chang MS, Kim WH, Lee BL. A hypoxia-dependent upregulation of hypoxia-inducible factor-1 by nuclear factor-kappaB promotes gastric tumour growth and angiogenesis. Br J Cancer. 2011;104:166–174. doi: 10.1038/sj.bjc.6606020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Yeramian A, Santacana M, Sorolla A, Llobet D, Encinas M, Velasco A, Bahi N, Eritja N, Domingo M, Oliva E, Dolcet X, Matias-Guiu X. Nuclear factor-kappaB2/p100 promotes endometrial carcinoma cell survival under hypoxia in a HIF-1alpha independent manner. Lab Invest. 2011;91:859–871. doi: 10.1038/labinvest.2011.58. [DOI] [PubMed] [Google Scholar]
  • 248.Cummins EP, Berra E, Comerford KM, Ginouves A, Fitzgerald KT, Seeballuck F, Godson C, Nielsen JE, Moynagh P, Pouyssegur J, Taylor CT. Prolyl hydroxylase-1 negatively regulates IkappaB kinase-beta, giving insight into hypoxia-induced NFkappaB activity. Proc Natl Acad Sci U S A. 2006;103:18154–18159. doi: 10.1073/pnas.0602235103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Melvin A, Mudie S, Rocha S. Further insights into the mechanism of hypoxia-induced NFkappaB. Cell Cycle. 2011;10:879–882. doi: 10.4161/cc.10.6.14910. [corrected] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Oliver KM, Taylor CT, Cummins EP. Hypoxia. Regulation of NFkappaB signalling during inflammation: the role of hydroxylases. Arthritis Res Ther. 2009;11:215. doi: 10.1186/ar2575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Ridder DA, Schwaninger M. NF-kappaB signaling in cerebral ischemia. Neuroscience. 2009;158:995–1006. doi: 10.1016/j.neuroscience.2008.07.007. [DOI] [PubMed] [Google Scholar]
  • 252.Rius J, Guma M, Schachtrup C, Akassoglou K, Zinkernagel AS, Nizet V, Johnson RS, Haddad GG, Karin M. NF-kappaB links innate immunity to the hypoxic response through transcriptional regulation of HIF-1alpha. Nature. 2008;453:807–811. doi: 10.1038/nature06905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Xue J, Li X, Jiao S, Wei Y, Wu G, Fang J. Prolyl hydroxylase-3 is downregulated in colorectal cancer cells and inhibits IKKbeta independent of hydroxylase activity. Gastroenterology. 2010;138:606–615. doi: 10.1053/j.gastro.2009.09.049. [DOI] [PubMed] [Google Scholar]
  • 254.Cockman ME, Lancaster DE, Stolze IP, Hewitson KS, McDonough MA, Coleman ML, Coles CH, Yu X, Hay RT, Ley SC, Pugh CW, Oldham NJ, Masson N, Schofield CJ, Ratcliffe PJ. Posttranslational hydroxylation of ankyrin repeats in IkappaB proteins by the hypoxia-inducible factor (HIF) asparaginyl hydroxylase, factor inhibiting HIF (FIH) Proc Natl Acad Sci U S A. 2006;103:14767–14772. doi: 10.1073/pnas.0606877103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Devries IL, Hampton-Smith RJ, Mulvihill MM, Alverdi V, Peet DJ, Komives EA. Consequences of IkappaB alpha hydroxylation by the factor inhibiting HIF (FIH) FEBS Lett. 2010;584:4725–4730. doi: 10.1016/j.febslet.2010.10.060. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 256.Shin DH, Li SH, Yang SW, Lee BL, Lee MK, Park JW. Inhibitor of nuclear factor-kappaB alpha derepresses hypoxia-inducible factor-1 during moderate hypoxia by sequestering factor inhibiting hypoxia-inducible factor from hypoxia-inducible factor 1alpha. FEBS J. 2009;276:3470–3480. doi: 10.1111/j.1742-4658.2009.07069.x. [DOI] [PubMed] [Google Scholar]
  • 257.Belaiba RS, Bonello S, Zahringer C, Schmidt S, Hess J, Kietzmann T, Gorlach A. Hypoxia up-regulates hypoxia-inducible factor-1alpha transcription by involving phosphatidylinositol 3-kinase and nuclear factor kappaB in pulmonary artery smooth muscle cells. Mol Biol Cell. 2007;18:4691–4697. doi: 10.1091/mbc.E07-04-0391. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Bonello S, Zahringer C, Belaiba RS, Djordjevic T, Hess J, Michiels C, Kietzmann T, Gorlach A. Reactive oxygen species activate the HIF-1alpha promoter via a functional NFkappaB site. Arterioscler Thromb Vasc Biol. 2007;27:755–761. doi: 10.1161/01.ATV.0000258979.92828.bc. [DOI] [PubMed] [Google Scholar]
  • 259.van Uden P, Kenneth NS, Rocha S. Regulation of hypoxia-inducible factor-1alpha by NF-kappaB. Biochem J. 2008;412:477–484. doi: 10.1042/BJ20080476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 260.Jiang H, Zhu YS, Xu H, Sun Y, Li QF. Inflammatory stimulation and hypoxia cooperatively activate HIF-1{alpha} in bronchial epithelial cells: involvement of PI3K and NF-{kappa}B. Am J Physiol Lung Cell Mol Physiol. 2010;298:L660–L669. doi: 10.1152/ajplung.00394.2009. [DOI] [PubMed] [Google Scholar]
  • 261.Lee SH, Lee YJ, Han HJ. Effect of arachidonic acid on hypoxia-induced IL-6 production in mouse ES cells: involvement of MAPKs, NF-kappaB, and HIF-1alpha. J Cell Physiol. 2010;222:574–585. doi: 10.1002/jcp.21973. [DOI] [PubMed] [Google Scholar]
  • 262.Trebec-Reynolds DP, Voronov I, Heersche JN, Manolson MF. VEGF-A expression in osteoclasts is regulated by NF-kappaB induction of HIF-1alpha. J Cell Biochem. 2010;110:343–351. doi: 10.1002/jcb.22542. [DOI] [PubMed] [Google Scholar]
  • 263.Kuo HP, Lee DF, Xia W, Wei Y, Hung MC. TNFalpha induces HIF-1alpha expression through activation of IKKbeta. Biochem Biophys Res Commun. 2009;389:640–644. doi: 10.1016/j.bbrc.2009.09.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Bracken CP, Whitelaw ML, Peet DJ. Activity of hypoxia-inducible factor 2alpha is regulated by association with the NF-kappaB essential modulator. J Biol Chem. 2005;280:14240–14251. doi: 10.1074/jbc.M409987200. [DOI] [PubMed] [Google Scholar]
  • 265.van Uden P, Kenneth NS, Webster R, Muller HA, Mudie S, Rocha S. Evolutionary conserved regulation of HIF-1beta by NF-kappaB. PLoS Genet. 2011;7:e1001285. doi: 10.1371/journal.pgen.1001285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 266.Bendinelli P, Matteucci E, Maroni P, Desiderio MA. NF-kappaB activation, dependent on acetylation/deacetylation, contributes to HIF-1 activity and migration of bone metastatic breast carcinoma cells. Mol Cancer Res. 2009;7:1328–1341. doi: 10.1158/1541-7786.MCR-08-0548. [DOI] [PubMed] [Google Scholar]
  • 267.Bader AG, Kang S, Zhao L, Vogt PK. Oncogenic PI3K deregulates transcription and translation. Nat Rev Cancer. 2005;5:921–929. doi: 10.1038/nrc1753. [DOI] [PubMed] [Google Scholar]
  • 268.Karar J, Maity A. PI3K/AKT/mTOR pathway in angiogenesis. Front Mol Neurosci. 2011;4:51. doi: 10.3389/fnmol.2011.00051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 269.Toker A, Cantley LC. Signalling through the lipid products of phosphoinositide-3-OH kinase. Nature. 1997;387:673–676. doi: 10.1038/42648. [DOI] [PubMed] [Google Scholar]
  • 270.Whitman M, Downes CP, Keeler M, Keller T, Cantley L. Type I phosphatidylinositol kinase makes a novel inositol phospholipid, phosphatidylinositol-3-phosphate. Nature. 1988;332:644–646. doi: 10.1038/332644a0. [DOI] [PubMed] [Google Scholar]
  • 271.Hollander MC, Blumenthal GM, Dennis PA. PTEN loss in the continuum of common cancers, rare syndromes and mouse models. Nat Rev Cancer. 2011;11:289–301. doi: 10.1038/nrc3037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Jiang BH, Liu LZ. PI3K/PTEN signaling in angiogenesis and tumorigenesis. Adv Cancer Res. 2009;102:19–65. doi: 10.1016/S0065-230X(09)02002-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Martelli AM, Evangelisti C, Follo MY, Ramazzotti G, Fini M, Giardino R, Manzoli L, McCubrey JA, Cocco L. Targeting the phosphatidylinositol 3-kinase/Akt/mammalian target of rapamycin signaling network in cancer stem cells. Curr Med Chem. 2011;18:2715–2726. doi: 10.2174/092986711796011201. [DOI] [PubMed] [Google Scholar]
  • 274.Sabbah M, Emami S, Redeuilh G, Julien S, Prevost G, Zimber A, Ouelaa R, Bracke M, De WO, Gespach C. Molecular signature and therapeutic perspective of the epithelial-to-mesenchymal transitions in epithelial cancers. Drug Resist Updat. 2008;11:123–151. doi: 10.1016/j.drup.2008.07.001. [DOI] [PubMed] [Google Scholar]
  • 275.DeBerardinis RJ, Lum JJ, Hatzivassiliou G, Thompson CB. The biology of cancer: metabolic reprogramming fuels cell growth and proliferation. Cell Metab. 2008;7:11–20. doi: 10.1016/j.cmet.2007.10.002. [DOI] [PubMed] [Google Scholar]
  • 276.Lee SH, Lee MY, Han HJ. Short-period hypoxia increases mouse embryonic stem cell proliferation through cooperation of arachidonic acid and PI3K/Akt signalling pathways. Cell Prolif. 2008;41:230–247. doi: 10.1111/j.1365-2184.2008.00516.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Lee SH, Lee MY, Lee JH, Han HJ. A potential mechanism for short time exposure to hypoxia-induced DNA synthesis in primary cultured chicken hepatocytes: correlation between Ca2+/PKC/MAPKs and PI3K/Akt/mTOR. J Cell Biochem. 2008;104:1598–1611. doi: 10.1002/jcb.21657. [DOI] [PubMed] [Google Scholar]
  • 278.Lu DY, Liou HC, Tang CH, Fu WM. Hypoxia-induced iNOS expression inmicroglia is regulated by the PI3-kinase/Akt/mTOR signaling pathway and activation of hypoxia inducible factor-1alpha. Biochem Pharmacol. 2006;72:992–1000. doi: 10.1016/j.bcp.2006.06.038. [DOI] [PubMed] [Google Scholar]
  • 279.Park KY, Lee HJ, Jeong SJ, Lee HJ, Kim HS, Kim SH, Lim S, Kim HC, Lu J, Kim SH. 1,2,3,4,6-Penta-O-galloly-beta-d-glucose suppresses hypoxia-induced accumulation of hypoxia-inducible factor-1alpha and signaling in LNCaP prostate cancer cells. Biol Pharm Bull. 2010;33:1835–1840. doi: 10.1248/bpb.33.1835. [DOI] [PubMed] [Google Scholar]
  • 280.Mottet D, Dumont V, Deccache Y, Demazy C, Ninane N, Raes M, Michiels C. Regulation of hypoxia-inducible factor-1alpha protein level during hypoxic conditions by the phosphatidylinositol 3-kinase/Akt/glycogen synthase kinase 3beta pathway in HepG2 cells. J Biol Chem. 2003;278:31277–31285. doi: 10.1074/jbc.M300763200. [DOI] [PubMed] [Google Scholar]
  • 281.Pore N, Jiang Z, Shu HK, Bernhard E, Kao GD, Maity A. Akt1 activation can augment hypoxia-inducible factor-1alpha expression by increasing protein translation through a mammalian target of rapamycin-independent pathway. Mol Cancer Res. 2006;4:471–479. doi: 10.1158/1541-7786.MCR-05-0234. [DOI] [PubMed] [Google Scholar]
  • 282.Laughner E, Taghavi P, Chiles K, Mahon PC, Semenza GL. HER2 (neu) signaling increases the rate of hypoxia-inducible factor 1alpha (HIF-1alpha) synthesis: novel mechanism for HIF-1-mediated vascular endothelial growth factor expression. Mol Cell Biol. 2001;21:3995–4004. doi: 10.1128/MCB.21.12.3995-4004.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.Sun HL, Liu YN, Huang YT, Pan SL, Huang DY, Guh JH, Lee FY, Kuo SC, Teng CM. YC-1 inhibits HIF-1 expression in prostate cancer cells: contribution of Akt/NF-kappaB signaling to HIF-1alpha accumulation during hypoxia. Oncogene. 2007;26:3941–3951. doi: 10.1038/sj.onc.1210169. [DOI] [PubMed] [Google Scholar]
  • 284.Treins C, Giorgetti-Peraldi S, Murdaca J, Semenza GL, Van OE. Insulin stimulates hypoxia-inducible factor 1 through a phosphatidylinositol 3-kinase/target of rapamycin-dependent signaling pathway. J Biol Chem. 2002;277:27975–27981. doi: 10.1074/jbc.M204152200. [DOI] [PubMed] [Google Scholar]
  • 285.Zhong H, Chiles K, Feldser D, Laughner E, Hanrahan C, Georgescu MM, Simons JW, Semenza GL. Modulation of hypoxia-inducible factor 1alpha expression by the epidermal growth factor/phosphatidylinositol 3-kinase/PTEN/AKT/FRAP pathway in human prostate cancer cells: implications for tumor angiogenesis and therapeutics. Cancer Res. 2000;60:1541–1545. [PubMed] [Google Scholar]
  • 286.Choi H, Chun YS, Kim TY, Park JW. HIF-2alpha enhances beta-catenin/TCF-driven transcription by interacting with beta-catenin. Cancer Res. 2010;70:10101–10111. doi: 10.1158/0008-5472.CAN-10-0505. [DOI] [PubMed] [Google Scholar]
  • 287.Sudhagar S, Sathya S, Lakshmi BS. Rapid non-genomic signalling by 17beta-oestradiol through c-Src involves mTOR-dependent expression of HIF-1alpha in breast cancer cells. Br J Cancer. 2011;105:953–960. doi: 10.1038/bjc.2011.349. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Villaume K, Blanc M, Gouysse G, Walter T, Couderc C, Nejjari M, Vercherat C, Cordier-Bussat M, Roche C, Scoazec JY. VEGF secretion by neuroendocrine tumor cells is inhibited by octreotide and by inhibitors of the PI3K/AKT/mTOR pathway. Neuroendocrinology. 2010;91:268–278. doi: 10.1159/000289569. [DOI] [PubMed] [Google Scholar]
  • 289.Zundel W, Schindler C, Haas-Kogan D, Koong A, Kaper F, Chen E, Gottschalk AR, Ryan HE, Johnson RS, Jefferson AB, Stokoe D, Giaccia AJ. Loss of PTEN facilitates HIF-1-mediated gene expression. Genes Dev. 2000;14:391–396. [PMC free article] [PubMed] [Google Scholar]
  • 290.Salama S, az-Arrastia C, Patel D, Botting S, Hatch S. 2-Methoxyestradiol, an endogenous estrogen metabolite, sensitizes radioresistant MCF-7/FIR breast cancer cells through multiple mechanisms. Int J Radiat Oncol Biol Phys. 2011;80:231–239. doi: 10.1016/j.ijrobp.2010.10.080. [DOI] [PubMed] [Google Scholar]
  • 291.Lino MM, Merlo A, Boulay JL. Notch signaling in glioblastoma: a developmental drug target? BMC Med. 2010;8:72. doi: 10.1186/1741-7015-8-72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 292.Artavanis-Tsakonas S, Rand MD, Lake RJ. Notch signaling: cell fate control and signal integration in development. Science. 1999;284:770–776. doi: 10.1126/science.284.5415.770. [DOI] [PubMed] [Google Scholar]
  • 293.Bray SJ. Notch signalling: a simple pathway becomes complex. Nat Rev Mol Cell Biol. 2006;7:678–689. doi: 10.1038/nrm2009. [DOI] [PubMed] [Google Scholar]
  • 294.Wang Z, Li Y, Banerjee S, Sarkar FH. Exploitation of the Notch signaling pathway as a novel target for cancer therapy. Anticancer Res. 2008;28:3621–3630. [PubMed] [Google Scholar]
  • 295.Wang Z, Li Y, Sarkar FH. Signaling mechanism(s) of reactive oxygen species in epithelial–mesenchymal transition reminiscent of cancer stem cells in tumor progression. Curr Stem Cell Res Ther. 2010;5:74–80. doi: 10.2174/157488810790442813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 296.Miele L. Notch signaling. Clin Cancer Res. 2006;12:1074–1079. doi: 10.1158/1078-0432.CCR-05-2570. [DOI] [PubMed] [Google Scholar]
  • 297.Wang Z, Ahmad A, Li Y, Azmi AS, Miele L, Sarkar FH. Targeting notch to eradicate pancreatic cancer stem cells for cancer therapy. Anticancer Res. 2011;31:1105–1113. [PubMed] [Google Scholar]
  • 298.Ellisen LW, Bird J, West DC, Soreng AL, Reynolds TC, Smith SD, Sklar J. TAN-1, the human homolog of the Drosophila notch gene, is broken by chromosomal translocations in T lymphoblastic neoplasms. Cell. 1991;66:649–661. doi: 10.1016/0092-8674(91)90111-b. [DOI] [PubMed] [Google Scholar]
  • 299.Radtke F, Raj K. The role of Notch in tumorigenesis: oncogene or tumour suppressor? Nat Rev Cancer. 2003;3:756–767. doi: 10.1038/nrc1186. [DOI] [PubMed] [Google Scholar]
  • 300.Weng AP, Ferrando AA, Lee W, Morris JP, Silverman LB, Sanchez-Irizarry C, Blacklow SC, Look AT, Aster JC. Activating mutations of NOTCH1 in human T cell acute lymphoblastic leukemia. Science. 2004;306:269–271. doi: 10.1126/science.1102160. [DOI] [PubMed] [Google Scholar]
  • 301.Chen Y, De Marco MA, Graziani I, Gazdar AF, Strack PR, Miele L, Bocchetta M. Oxygen concentration determines the biological effects of NOTCH-1 signaling in adenocarcinoma of the lung. Cancer Res. 2007;67:7954–7959. doi: 10.1158/0008-5472.CAN-07-1229. [DOI] [PubMed] [Google Scholar]
  • 302.Santagata S, Demichelis F, Riva A, Varambally S, Hofer MD, Kutok JL, Kim R, Tang J, Montie JE, Chinnaiyan AM, Rubin MA, Aster JC. JAGGED1 expression is associated with prostate cancer metastasis and recurrence. Cancer Res. 2004;64:6854–6857. doi: 10.1158/0008-5472.CAN-04-2500. [DOI] [PubMed] [Google Scholar]
  • 303.Soares R, Balogh G, Guo S, Gartner F, Russo J, Schmitt F. Evidence for the notch signaling pathway on the role of estrogen in angiogenesis. Mol Endocrinol. 2004;18:2333–2343. doi: 10.1210/me.2003-0362. [DOI] [PubMed] [Google Scholar]
  • 304.Qiang L, Wu T, Zhang HW, Lu N, Hu R, Wang YJ, Zhao L, Chen FH, Wang XT, You QD, Guo QL. HIF-1alpha is critical for hypoxia-mediated maintenance of glioblastoma stem cells by activating Notch signaling pathway. Cell Death Differ. 2012;19:284–294. doi: 10.1038/cdd.2011.95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 305.Bedogni B, Warneke JA, Nickoloff BJ, Giaccia AJ, Powell MB. Notch1 is an effector of Akt and hypoxia in melanoma development. J Clin Invest. 2008;118:3660–3670. doi: 10.1172/JCI36157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 306.Guo S, Liu M, Gonzalez-Perez RR. Role of Notch and its oncogenic signaling crosstalk in breast cancer. Biochim Biophys Acta. 2011;1815:197–213. doi: 10.1016/j.bbcan.2010.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Mukherjee T, Kim WS, Mandal L, Banerjee U. Interaction between Notch and Hif-alpha in development and survival of Drosophila blood cells. Science. 2011;332:1210–1213. doi: 10.1126/science.1199643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 308.Eliasz S, Liang S, Chen Y, De Marco MA, Machek O, Skucha S, Miele L, Bocchetta M. Notch-1 stimulates survival of lung adenocarcinoma cells during hypoxia by activating the IGF-1R pathway. Oncogene. 2010;29:2488–2498. doi: 10.1038/onc.2010.7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 309.Pistollato F, Rampazzo E, Persano L, Abbadi S, Frasson C, Denaro L, D’Avella D, Panchision DM, Della PA, Scienza R, Basso G. Interaction of hypoxia-inducible factor-1alpha and Notch signaling regulates medulloblastoma precursor proliferation and fate. Stem Cells. 2010;28:1918–1929. doi: 10.1002/stem.518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 310.Lee SH, Kim MH, Han HJ. Arachidonic acid potentiates hypoxia-induced VEGF expression in mouse embryonic stem cells: involvement of Notch, Wnt, and HIF-1alpha. Am J Physiol Cell Physiol. 2009;297:C207–C216. doi: 10.1152/ajpcell.00579.2008. [DOI] [PubMed] [Google Scholar]
  • 311.Coleman ML, McDonough MA, Hewitson KS, Coles C, Mecinovic J, Edelmann M, Cook KM, Cockman ME, Lancaster DE, Kessler BM, Oldham NJ, Ratcliffe PJ, Schofield CJ. Asparaginyl hydroxylation of the Notch ankyrin repeat domain by factor inhibiting hypoxia-inducible factor. J Biol Chem. 2007;282:24027–24038. doi: 10.1074/jbc.M704102200. [DOI] [PubMed] [Google Scholar]
  • 312.Zheng X, Linke S, Dias JM, Zheng X, Gradin K, Wallis TP, Hamilton BR, Gustafsson M, Ruas JL, Wilkins S, Bilton RL, Brismar K, Whitelaw ML, Pereira T, Gorman JJ, Ericson J, Peet DJ, Lendahl U, Poellinger L. Interaction with factor inhibiting HIF-1 defines an additional mode of cross-coupling between the Notch and hypoxia signaling pathways. Proc Natl Acad Sci U S A. 2008;105:3368–3373. doi: 10.1073/pnas.0711591105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 313.Poellinger L, Lendahl U. Modulating Notch signaling by pathway-intrinsic and pathway-extrinsic mechanisms. Curr Opin Genet Dev. 2008;18:449–454. doi: 10.1016/j.gde.2008.07.013. [DOI] [PubMed] [Google Scholar]
  • 314.DeSano JT, Xu L. MicroRNA regulation of cancer stem cells and therapeutic implications. AAPS J. 2009;11:682–692. doi: 10.1208/s12248-009-9147-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Perera RJ, Ray A. MicroRNAs in the search for understanding human diseases. BioDrugs. 2007;21:97–104. doi: 10.2165/00063030-200721020-00004. [DOI] [PubMed] [Google Scholar]
  • 316.Bienz M, Clevers H. Armadillo/beta-catenin signals in the nucleus—proof beyond a reasonable doubt? Nat Cell Biol. 2003;5:179–182. doi: 10.1038/ncb0303-179. [DOI] [PubMed] [Google Scholar]
  • 317.Peifer M, Polakis P. Wnt signaling in oncogenesis and embryogenesis—a look outside the nucleus. Science. 2000;287:1606–1609. doi: 10.1126/science.287.5458.1606. [DOI] [PubMed] [Google Scholar]
  • 318.Doucas H, Garcea G, Neal CP, Manson MM, Berry DP. Changes in the Wnt signalling pathway in gastrointestinal cancers and their prognostic significance. Eur J Cancer. 2005;41:365–379. doi: 10.1016/j.ejca.2004.11.005. [DOI] [PubMed] [Google Scholar]
  • 319.Kang CM, Kim HK, Kim H, Choi GH, Kim KS, Choi JS, Lee WJ. Expression of Wnt target genes in solid pseudopapillary tumor of the pancreas: a pilot study. Pancreas. 2009;38:e53–e59. doi: 10.1097/mpa.0b013e318196d198. [DOI] [PubMed] [Google Scholar]
  • 320.Honeycutt KA, Roop DR. c-Myc and epidermal stem cell fate determination. J Dermatol. 2004;31:368–375. doi: 10.1111/j.1346-8138.2004.tb00687.x. [DOI] [PubMed] [Google Scholar]
  • 321.Honeycutt KA, Koster MI, Roop DR. Genes involved in stem cell fate decisions and commitment to differentiation play a role in skin disease. J Investig Dermatol Symp Proc. 2004;9:261–268. doi: 10.1111/j.1087-0024.2004.09312.x. [DOI] [PubMed] [Google Scholar]
  • 322.Takebe N, Harris PJ, Warren RQ, Ivy SP. Targeting cancer stem cells by inhibiting Wnt, Notch, and Hedgehog pathways. Nat Rev Clin Oncol. 2011;8:97–106. doi: 10.1038/nrclinonc.2010.196. [DOI] [PubMed] [Google Scholar]
  • 323.Katoh M, Katoh M. WNT signaling pathway and stem cell signaling network. Clin Cancer Res. 2007;13:4042–4045. doi: 10.1158/1078-0432.CCR-06-2316. [DOI] [PubMed] [Google Scholar]
  • 324.Jiang YG, Luo Y, He DL, Li X, Zhang LL, Peng T, Li MC, Lin YH. Role of Wnt/beta-catenin signaling pathway in epithelial–mesenchymal transition of human prostate cancer induced by hypoxia-inducible factor-1alpha. Int J Urol. 2007;14:1034–1039. doi: 10.1111/j.1442-2042.2007.01866.x. [DOI] [PubMed] [Google Scholar]
  • 325.Luo Y, He DL, Ning L, Shen SL, Li L, Li X, Zhau HE, Chung LW. Over-expression of hypoxia-inducible factor-1alpha increases the invasive potency of LNCaP cells in vitro. BJU Int. 2006;98:1315–1319. doi: 10.1111/j.1464-410X.2006.06480.x. [DOI] [PubMed] [Google Scholar]
  • 326.Liu S, Dontu G, Wicha MS. Mammary stem cells, self-renewal pathways, and carcinogenesis. Breast Cancer Res. 2005;7:86–95. doi: 10.1186/bcr1021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 327.de Sousa EM, Vermeulen L, Richel D, Medema JP. Targeting Wnt signaling in colon cancer stem cells. Clin Cancer Res. 2011;17:647–653. doi: 10.1158/1078-0432.CCR-10-1204. [DOI] [PubMed] [Google Scholar]
  • 328.Katoh M. Network of WNT and other regulatory signaling cascades in pluripotent stem cells and cancer stem cells. Curr Pharm Biotechnol. 2011;12:160–170. doi: 10.2174/138920111794295710. [DOI] [PubMed] [Google Scholar]
  • 329.Oishi N, Wang XW. Novel therapeutic strategies for targeting liver cancer stem cells. Int J Biol Sci. 2011;7:517–535. doi: 10.7150/ijbs.7.517. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 330.Kaidi A, Williams AC, Paraskeva C. Interaction between beta-catenin and HIF-1 promotes cellular adaptation to hypoxia. Nat Cell Biol. 2007;9:210–217. doi: 10.1038/ncb1534. [DOI] [PubMed] [Google Scholar]
  • 331.Lim JH, Chun YS, Park JW. Hypoxia-inducible factor-1alpha obstructs a Wnt signaling pathway by inhibiting the hARD1-mediated activation of beta-catenin. Cancer Res. 2008;68:5177–5184. doi: 10.1158/0008-5472.CAN-07-6234. [DOI] [PubMed] [Google Scholar]
  • 332.Genetos DC, Toupadakis CA, Raheja LF, Wong A, Papanicolaou SE, Fyhrie DP, Loots GG, Yellowley CE. Hypoxia decreases sclerostin expression and increases Wnt signaling in osteoblasts. J Cell Biochem. 2010;110:457–467. doi: 10.1002/jcb.22559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333.Giles RH, Lolkema MP, Snijckers CM, Belderbos M, van der GP, Mans DA, van BM, van NM, Goldschmeding R, van Diest PJ, Clevers H, Voest EE. Interplay between VHL/HIF1alpha and Wnt/beta-catenin pathways during colorectal tumorigenesis. Oncogene. 2006;25:3065–3070. doi: 10.1038/sj.onc.1209330. [DOI] [PubMed] [Google Scholar]
  • 334.Zhao JH, Luo Y, Jiang YG, He DL, Wu CT. Knockdown of beta-catenin through shRNA cause a reversal of EMT and metastatic phenotypes induced by HIF-1alpha. Cancer Invest. 2011;29:377–382. doi: 10.3109/07357907.2010.512595. [DOI] [PubMed] [Google Scholar]
  • 335.Mazumdar J, O’Brien WT, Johnson RS, LaManna JC, Chavez JC, Klein PS, Simon MC. O2 regulates stem cells through Wnt/beta-catenin signalling. Nat Cell Biol. 2010;12:1007–1013. doi: 10.1038/ncb2102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 336.Cui XP, Xing Y, Chen JM, Dong SW, Ying DJ, Yew DT. Wnt/beta-catenin is involved in the proliferation of hippocampal neural stem cells induced by hypoxia. Ir J Med Sci. 2011;180:387–393. doi: 10.1007/s11845-010-0566-3. [DOI] [PubMed] [Google Scholar]
  • 337.Gritli-Linde A, Bei M, Maas R, Zhang XM, Linde A, McMahon AP. Shh signaling within the dental epithelium is necessary for cell proliferation, growth and polarization. Development. 2002;129:5323–5337. doi: 10.1242/dev.00100. [DOI] [PubMed] [Google Scholar]
  • 338.Yang L, Xie G, Fan Q, Xie J. Activation of the hedgehog-signaling pathway in human cancer and the clinical implications. Oncogene. 2010;29:469–481. doi: 10.1038/onc.2009.392. [DOI] [PubMed] [Google Scholar]
  • 339.Chaudary N, Pintilie M, Hedley D, Fyles AW, Milosevic M, Clarke B, Hill RP, Mackay H. Hedgehog pathway signaling in cervical carcinoma and outcome after chemoradiation. Cancer. 2012;118:3105–3115. doi: 10.1002/cncr.26635. [DOI] [PubMed] [Google Scholar]
  • 340.Dormoy V, Danilin S, Lindner V, Thomas L, Rothhut S, Coquard C, Helwig JJ, Jacqmin D, Lang H, Massfelder T. The sonic hedgehog signaling pathway is reactivated in human renal cell carcinoma and plays orchestral role in tumor growth. Mol Cancer. 2009;8:123. doi: 10.1186/1476-4598-8-123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 341.Choi SS, Omenetti A, Witek RP, Moylan CA, Syn WK, Jung Y, Yang L, Sudan DL, Sicklick JK, Michelotti GA, Rojkind M, Diehl AM. Hedgehog pathway activation and epithelial-to-mesenchymal transitions during myofibroblastic transformation of rat hepatic cells in culture and cirrhosis. Am J Physiol Gastrointest Liver Physiol. 2009;297:G1093–G1106. doi: 10.1152/ajpgi.00292.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 342.Sarkar FH, Li Y, Wang Z, Kong D. The role of nutraceuticals in the regulation of Wnt and Hedgehog signaling in cancer. Cancer Metastasis Rev. 2010;29:383–394. doi: 10.1007/s10555-010-9233-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 343.Bicknell R, Harris AL. Novel angiogenic signaling pathways and vascular targets. Annu Rev Pharmacol Toxicol. 2004;44:219–238. doi: 10.1146/annurev.pharmtox.44.101802.121650. [DOI] [PubMed] [Google Scholar]
  • 344.Pola R, Ling LE, Silver M, Corbley MJ, Kearney M, Blake PR, Shapiro R, Taylor FR, Baker DP, Asahara T, Isner JM. The morphogen Sonic hedgehog is an indirect angiogenic agent upregulating two families of angiogenic growth factors. Nat Med. 2001;7:706–711. doi: 10.1038/89083. [DOI] [PubMed] [Google Scholar]
  • 345.Lawson ND, Weinstein BM. Arteries and veins: making a difference with zebrafish. Nat Rev Genet. 2002;3:674–682. doi: 10.1038/nrg888. [DOI] [PubMed] [Google Scholar]
  • 346.Lawson ND, Vogel AM, Weinstein BM. Sonic hedgehog and vascular endothelial growth factor act upstream of the Notch pathway during arterial endothelial differentiation. Dev Cell. 2002;3:127–136. doi: 10.1016/s1534-5807(02)00198-3. [DOI] [PubMed] [Google Scholar]
  • 347.Hwang JM, Weng YJ, Lin JA, Bau DT, Ko FY, Tsai FJ, Tsai CH, Wu CH, Lin PC, Huang CY, Kuo WW. Hypoxia-induced compensatory effect as related to Shh and HIF-1alpha in ischemia embryo rat heart. Mol Cell Biochem. 2008;311:179–187. doi: 10.1007/s11010-008-9708-6. [DOI] [PubMed] [Google Scholar]
  • 348.Wang G, Zhang Z, Xu Z, Yin H, Bai L, Ma Z, Decoster MA, Qian G, Wu G. Activation of the sonic hedgehog signaling controls human pulmonary arterial smooth muscle cell proliferation in response to hypoxia. Biochim Biophys Acta. 2010;1803:1359–1367. doi: 10.1016/j.bbamcr.2010.09.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 349.Onishi H, Kai M, Odate S, Iwasaki H, Morifuji Y, Ogino T, Morisaki T, Nakashima Y, Katano M. Hypoxia activates the hedgehog signaling pathway in a ligand-independent manner by upregulation of Smo transcription in pancreatic cancer. Cancer Sci. 2011;102:1144–1150. doi: 10.1111/j.1349-7006.2011.01912.x. [DOI] [PubMed] [Google Scholar]
  • 350.Garzon R, Fabbri M, Cimmino A, Calin GA, Croce CM. MicroRNA expression and function in cancer. Trends Mol Med. 2006;12:580–587. doi: 10.1016/j.molmed.2006.10.006. [DOI] [PubMed] [Google Scholar]
  • 351.Kulshreshtha R, Ferracin M, Negrini M, Calin GA, Davuluri RV, Ivan M. Regulation of microRNA expression: the hypoxic component. Cell Cycle. 2007;6:1426–1431. [PubMed] [Google Scholar]
  • 352.Kulshreshtha R, Ferracin M, Wojcik SE, Garzon R, Alder H, gosto-Perez FJ, Davuluri R, Liu CG, Croce CM, Negrini M, Calin GA, Ivan M. A microRNA signature of hypoxia. Mol Cell Biol. 2007;27:1859–1867. doi: 10.1128/MCB.01395-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 353.Kulshreshtha R, Davuluri RV, Calin GA, Ivan M. A microRNA component of the hypoxic response. Cell Death Differ. 2008;15:667–671. doi: 10.1038/sj.cdd.4402310. [DOI] [PubMed] [Google Scholar]
  • 354.Pocock R. Invited review: decoding the microRNA response to hypoxia. Pflugers Arch. 2011;461:307–315. doi: 10.1007/s00424-010-0910-5. [DOI] [PubMed] [Google Scholar]
  • 355.Dillhoff M, Liu J, Frankel W, Croce C, Bloomston M. MicroRNA-21 is overexpressed in pancreatic cancer and a potential predictor of survival. J Gastrointest Surg. 2008;12:2171–2176. doi: 10.1007/s11605-008-0584-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 356.Moriyama T, Ohuchida K, Mizumoto K, Yu J, Sato N, Nabae T, Takahata S, Toma H, Nagai E, Tanaka M. MicroRNA-21 modulates biological functions of pancreatic cancer cells including their proliferation, invasion, and chemoresistance. Mol Cancer Ther. 2009;8:1067–1074. doi: 10.1158/1535-7163.MCT-08-0592. [DOI] [PubMed] [Google Scholar]
  • 357.Ali S, Ahmad A, Banerjee S, Padhye S, Dominiak K, Schaffert JM, Wang Z, Philip PA, Sarkar FH. Gemcitabine sensitivity can be induced in pancreatic cancer cells through modulation of miR-200 and miR-21 expression by curcumin or its analogue CDF. Cancer Res. 2010;70:3606–3617. doi: 10.1158/0008-5472.CAN-09-4598. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 358.Bao B, Ali S, Banerjee S, Wang Z, Logna F, Azmi AS, Kong D, Ahmad A, Li Y, Padhye S, Sarkar FH. Curcumin analogue CDF inhibits pancreatic tumor growth by switching on suppressor microRNAs and attenuating EZH2 expression. Cancer Res. 2012;72:335–345. doi: 10.1158/0008-5472.CAN-11-2182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 359.Olson P, Lu J, Zhang H, Shai A, Chun MG, Wang Y, Libutti SK, Nakakura EK, Golub TR, Hanahan D. MicroRNA dynamics in the stages of tumorigenesis correlate with hallmark capabilities of cancer. Genes Dev. 2009;23:2152–2165. doi: 10.1101/gad.1820109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 360.Zhang B, Pan X, Cobb GP, Anderson TA. microRNAs as oncogenes and tumor suppressors. Dev Biol. 2007;302:1–12. doi: 10.1016/j.ydbio.2006.08.028. [DOI] [PubMed] [Google Scholar]
  • 361.Liu LZ, Li C, Chen Q, Jing Y, Carpenter R, Jiang Y, Kung HF, Lai L, Jiang BH. MiR-21 induced angiogenesis through AKT and ERK activation and HIF-1alpha expression. PLoS One. 2011;6:e19139. doi: 10.1371/journal.pone.0019139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 362.Nie Y, Han BM, Liu XB, Yang JJ, Wang F, Cong XF, Chen X. Identification of microRNAs involved in hypoxia- and serum deprivation-induced apoptosis in mesenchymal stem cells. Int J Biol Sci. 2011;7:762–768. doi: 10.7150/ijbs.7.762. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 363.Yamakuchi M, Lotterman CD, Bao C, Hruban RH, Karim B, Mendell JT, Huso D, Lowenstein CJ. P53-induced microRNA-107 inhibits HIF-1 and tumor angiogenesis. Proc Natl Acad Sci U S A. 2010;107:6334–6339. doi: 10.1073/pnas.0911082107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 364.Jiang J, Zheng X, Xu X, Zhou Q, Yan H, Zhang X, Lu B, Wu C, Ju J. Prognostic significance of miR-181b and miR-21 in gastric cancer patients treated with S-1/Oxaliplatin or Doxifluridine/Oxaliplatin. PLoS One. 2011;6:e23271. doi: 10.1371/journal.pone.0023271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 365.Visone R, Veronese A, Rassenti LZ, Balatti V, Pearl DK, Acunzo M, Volinia S, Taccioli C, Kipps TJ, Croce CM. miR-181b is a biomarker of disease progression in chronic lymphocytic leukemia. Blood. 2011;118:3072–3079. doi: 10.1182/blood-2011-01-333484. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 366.Zhi F, Chen X, Wang S, Xia X, Shi Y, Guan W, Shao N, Qu H, Yang C, Zhang Y, Wang Q, Wang R, Zen K, Zhang CY, Zhang J, Yang Y. The use of hsa-miR-21, hsa-miR-181b and hsa-miR-106a as prognostic indicators of astrocytoma. Eur J Cancer. 2010;46:1640–1649. doi: 10.1016/j.ejca.2010.02.003. [DOI] [PubMed] [Google Scholar]
  • 367.Ciafre SA, Galardi S, Mangiola A, Ferracin M, Liu CG, Sabatino G, Negrini M, Maira G, Croce CM, Farace MG. Extensive modulation of a set of microRNAs in primary glioblastoma. Biochem Biophys Res Commun. 2005;334:1351–1358. doi: 10.1016/j.bbrc.2005.07.030. [DOI] [PubMed] [Google Scholar]
  • 368.Garzon R, Pichiorri F, Palumbo T, Visentini M, Aqeilan R, Cimmino A, Wang H, Sun H, Volinia S, Alder H, Calin GA, Liu CG, Andreeff M, Croce CM. MicroRNA gene expression during retinoic acid-induced differentiation of human acute promyelocytic leukemia. Oncogene. 2007;26:4148–4157. doi: 10.1038/sj.onc.1210186. [DOI] [PubMed] [Google Scholar]
  • 369.Zanette DL, Rivadavia F, Molfetta GA, Barbuzano FG, Proto-Siqueira R, Silva WA, Jr, Falcao RP, Zago MA. miRNA expression profiles in chronic lymphocytic and acute lymphocytic leukemia. Braz J Med Biol Res. 2007;40:1435–1440. doi: 10.1590/s0100-879x2007001100003. [DOI] [PubMed] [Google Scholar]
  • 370.Chen H, Chen Q, Fang M, Mi Y. microRNA-181b targets MLK2 in HL-60 cells. Sci China Life Sci. 2010;53:101–106. doi: 10.1007/s11427-010-0002-y. [DOI] [PubMed] [Google Scholar]
  • 371.Zhu W, Shan X, Wang T, Shu Y, Liu P. miR-181b modulates multidrug resistance by targeting BCL2 in human cancer cell lines. Int J Cancer. 2010;127:2520–2529. doi: 10.1002/ijc.25260. [DOI] [PubMed] [Google Scholar]
  • 372.Wang B, Hsu SH, Majumder S, Kutay H, Huang W, Jacob ST, Ghoshal K. TGFbeta-mediated upregulation of hepatic miR-181b promotes hepatocarcinogenesis by targeting TIMP3. Oncogene. 2010;29:1787–1797. doi: 10.1038/onc.2009.468. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 373.Shi L, Cheng Z, Zhang J, Li R, Zhao P, Fu Z, You Y. hsa-mir-181a and hsa-mir-181b function as tumor suppressors in human glioma cells. Brain Res. 2008;1236:185–193. doi: 10.1016/j.brainres.2008.07.085. [DOI] [PubMed] [Google Scholar]
  • 374.Hebert C, Norris K, Scheper MA, Nikitakis N, Sauk JJ. High mobility group A2 is a target for miRNA-98 in head and neck squamous cell carcinoma. Mol Cancer. 2007;6:5. doi: 10.1186/1476-4598-6-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 375.Xu X, Jia R, Zhou Y, Song X, Wang J, Qian G, Ge S, Fan X. Microarray-based analysis: identification of hypoxia-regulated microRNAs in retinoblastoma cells. Int J Oncol. 2011;38:1385–1393. doi: 10.3892/ijo.2011.961. [DOI] [PubMed] [Google Scholar]
  • 376.Camps C, Buffa FM, Colella S, Moore J, Sotiriou C, Sheldon H, Harris AL, Gleadle JM, Ragoussis J. hsa-miR-210 is induced by hypoxia and is an independent prognostic factor in breast cancer. Clin Cancer Res. 2008;14:1340–1348. doi: 10.1158/1078-0432.CCR-07-1755. [DOI] [PubMed] [Google Scholar]
  • 377.Chan SY, Loscalzo J. MicroRNA-210: a unique and pleiotropic hypoxamir. Cell Cycle. 2010;9:1072–1083. doi: 10.4161/cc.9.6.11006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 378.Devlin C, Greco S, Martelli F, Ivan M. miR-210: More than a silent player in hypoxia. IUBMB Life. 2011;63:94–100. doi: 10.1002/iub.427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 379.Favaro E, Ramachandran A, McCormick R, Gee H, Blancher C, Crosby M, Devlin C, Blick C, Buffa F, Li JL, Vojnovic B, Pires das NR, Glazer P, Iborra F, Ivan M, Ragoussis J, Harris AL. MicroRNA-210 regulates mitochondrial free radical response to hypoxia and krebs cycle in cancer cells by targeting iron sulfur cluster protein ISCU. PLoS One. 2010;5:e10345. doi: 10.1371/journal.pone.0010345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 380.Gee HE, Camps C, Buffa FM, Patiar S, Winter SC, Betts G, Homer J, Corbridge R, Cox G, West CM, Ragoussis J, Harris AL. hsa-mir-210 is a marker of tumor hypoxia and a prognostic factor in head and neck cancer. Cancer. 2010;116:2148–2158. doi: 10.1002/cncr.25009. [DOI] [PubMed] [Google Scholar]
  • 381.Ho AS, Huang X, Cao H, Christman-Skieller C, Bennewith K, Le QT, Koong AC. Circulating miR-210 as a novel hypoxia marker in pancreatic cancer. Transl Oncol. 2010;3:109–113. doi: 10.1593/tlo.09256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 382.Puissegur MP, Mazure NM, Bertero T, Pradelli L, Grosso S, Robbe-Sermesant K, Maurin T, Lebrigand K, Cardinaud B, Hofman V, Fourre S, Magnone V, Ricci JE, Pouyssegur J, Gounon P, Hofman P, Barbry P, Mari B. miR-210 is overexpressed in late stages of lung cancer and mediates mitochondrial alterations associated with modulation of HIF-1 activity. Cell Death Differ. 2011;18:465–478. doi: 10.1038/cdd.2010.119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 383.Quero L, Dubois L, Lieuwes NG, Hennequin C, Lambin P. miR-210 as a marker of chronic hypoxia, but not a therapeutic target in prostate cancer. Radiother Oncol. 2011;101:203–208. doi: 10.1016/j.radonc.2011.05.063. [DOI] [PubMed] [Google Scholar]
  • 384.Huang X, Ding L, Bennewith KL, Tong RT, Welford SM, Ang KK, Story M, Le QT, Giaccia AJ. Hypoxia-inducible mir-210 regulates normoxic gene expression involved in tumor initiation. Mol Cell. 2009;35:856–867. doi: 10.1016/j.molcel.2009.09.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 385.Ivan M, Harris AL, Martelli F, Kulshreshtha R. Hypoxia response and microRNAs: no longer two separate worlds. J Cell Mol Med. 2008;12:1426–1431. doi: 10.1111/j.1582-4934.2008.00398.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 386.Hu S, Huang M, Li Z, Jia F, Ghosh Z, Lijkwan MA, Fasanaro P, Sun N, Wang X, Martelli F, Robbins RC, Wu JC. MicroRNA-210 as a novel therapy for treatment of ischemic heart disease. Circulation. 2010;122:S124–S131. doi: 10.1161/CIRCULATIONAHA.109.928424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 387.Hua Z, Lv Q, Ye W, Wong CK, Cai G, Gu D, Ji Y, Zhao C, Wang J, Yang BB, Zhang Y. MiRNA-directed regulation of VEGF and other angiogenic factors under hypoxia. PLoS One. 2006;1:e116. doi: 10.1371/journal.pone.0000116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 388.Huang X, Le QT, Giaccia AJ. MiR-210—micromanager of the hypoxia pathway. Trends Mol Med. 2010;16:230–237. doi: 10.1016/j.molmed.2010.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 389.Crosby ME, Kulshreshtha R, Ivan M, Glazer PM. MicroRNA regulation of DNA repair gene expression in hypoxic stress. Cancer Res. 2009;69:1221–1229. doi: 10.1158/0008-5472.CAN-08-2516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 390.Crosby ME, Devlin CM, Glazer PM, Calin GA, Ivan M. Emerging roles of microRNAs in the molecular responses to hypoxia. Curr Pharm Des. 2009;15:3861–3866. doi: 10.2174/138161209789649367. [DOI] [PubMed] [Google Scholar]
  • 391.Kong D, Banerjee S, Ahmad A, Li Y, Wang Z, Sethi S, Sarkar FH. Epithelial to mesenchymal transition is mechanistically linked with stem cell signatures in prostate cancer cells. PLoS One. 2010;5:e12445. doi: 10.1371/journal.pone.0012445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 392.Chang CJ, Hsu CC, Chang CH, Tsai LL, Chang YC, Lu SW, Yu CH, Huang HS, Wang JJ, Tsai CH, Chou MY, Yu CC, Hu FW. Let-7d functions as novel regulator of epithelial–mesenchymal transition and chemoresistant property in oral cancer. Oncol Rep. 2011;26:1003–1010. doi: 10.3892/or.2011.1360. [DOI] [PubMed] [Google Scholar]
  • 393.McCarty MF. Metformin may antagonize Lin28 and/or Lin28B activity, thereby boosting let-7 levels and antagonizing cancer progression. Med Hypotheses. 2012;78:262–269. doi: 10.1016/j.mehy.2011.10.041. [DOI] [PubMed] [Google Scholar]
  • 394.Li Y, VandenBoom TG, Kong D, Wang Z, Ali S, Philip PA, Sarkar FH. Up-regulation of miR-200 and let-7 by natural agents leads to the reversal of epithelial-to-mesenchymal transition in gemcitabine-resistant pancreatic cancer cells. Cancer Res. 2009;69:6704–6712. doi: 10.1158/0008-5472.CAN-09-1298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 395.Peter ME. Let-7 and miR-200 microRNAs: guardians against pluripotency and cancer progression. Cell Cycle. 2009;8:843–852. doi: 10.4161/cc.8.6.7907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 396.Cascio S, D’Andrea A, Ferla R, Surmacz E, Gulotta E, Amodeo V, Bazan V, Gebbia N, Russo A. miR-20b modulates VEGF expression by targeting HIF-1 alpha and STAT3 in MCF-7 breast cancer cells. J Cell Physiol. 2010;224:242–249. doi: 10.1002/jcp.22126. [DOI] [PubMed] [Google Scholar]
  • 397.Li J, Zhang Y, Zhao J, Kong F, Chen Y. Overexpression of miR-22 reverses paclitaxel-induced chemoresistance through activation of PTEN signaling in p53-mutated colon cancer cells. Mol Cell Biochem. 2011;357:31–38. doi: 10.1007/s11010-011-0872-8. [DOI] [PubMed] [Google Scholar]
  • 398.Tsuchiya N, Izumiya M, Ogata-Kawata H, Okamoto K, Fujiwara Y, Nakai M, Okabe A, Schetter AJ, Bowman ED, Midorikawa Y, Sugiyama Y, Aburatani H, Harris CC, Nakagama H. Tumor suppressor miR-22 determines p53-dependent cellular fate through post-transcriptional regulation of p21. Cancer Res. 2011;71:4628–4639. doi: 10.1158/0008-5472.CAN-10-2475. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 399.Li J, Liang S, Yu H, Zhang J, Ma D, Lu X. An inhibitory effect of miR-22 on cell migration and invasion in ovarian cancer. Gynecol Oncol. 2010;119:543–548. doi: 10.1016/j.ygyno.2010.08.034. [DOI] [PubMed] [Google Scholar]
  • 400.Zhang J, Yang Y, Yang T, Liu Y, Li A, Fu S, Wu M, Pan Z, Zhou W. microRNA-22, downregulated in hepatocellular carcinoma and correlated with prognosis, suppresses cell proliferation and tumourigenicity. Br J Cancer. 2010;103:1215–1220. doi: 10.1038/sj.bjc.6605895. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 401.Pandey DP, Picard D. miR-22 inhibits estrogen signaling by directly targeting the estrogen receptor alpha mRNA. Mol Cell Biol. 2009;29:3783–3790. doi: 10.1128/MCB.01875-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 402.Xiong J, Yu D, Wei N, Fu H, Cai T, Huang Y, Wu C, Zheng X, Du Q, Lin D, Liang Z. An estrogen receptor alpha suppressor, microRNA-22, is downregulated in estrogen receptor alpha-positive human breast cancer cell lines and clinical samples. FEBS J. 2010;277:1684–1694. doi: 10.1111/j.1742-4658.2010.07594.x. [DOI] [PubMed] [Google Scholar]
  • 403.Yamakuchi M, Yagi S, Ito T, Lowenstein CJ. MicroRNA-22 regulates hypoxia signaling in colon cancer cells. PLoS One. 2011;6:e20291. doi: 10.1371/journal.pone.0020291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 404.Sparmann A, van LM. Polycomb silencers control cell fate, development and cancer. Nat Rev Cancer. 2006;6:846–856. doi: 10.1038/nrc1991. [DOI] [PubMed] [Google Scholar]
  • 405.Cao P, Deng Z, Wan M, Huang W, Cramer SD, Xu J, Lei M, Sui G. MicroRNA-101 negatively regulates Ezh2 and its expression is modulated by androgen receptor and HIF-1alpha/HIF-1beta. Mol Cancer. 2010;9:108. doi: 10.1186/1476-4598-9-108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 406.Zhang H, Hao Y, Yang J, Zhou Y, Li J, Yin S, Sun C, Ma M, Huang Y, Xi JJ. Genome-wide functional screening of miR-23b as a pleiotropic modulator suppressing cancer metastasis. Nat Commun. 2011;2:554. doi: 10.1038/ncomms1555. [DOI] [PubMed] [Google Scholar]
  • 407.Sun T, Yang M, Chen S, Balk S, Pomerantz M, Hsieh CL, Brown M, Lee GS, Kantoff PW. The altered expression of MiR-221/-222 and MiR-23b/-27b is associated with the development of human castration resistant prostate cancer. Prostate. 2012;72:1093–1103. doi: 10.1002/pros.22456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 408.Liu W, Zabirnyk O, Wang H, Shiao YH, Nickerson ML, Khalil S, Anderson LM, Perantoni AO, Phang JM. miR-23b targets proline oxidase, a novel tumor suppressor protein in renal cancer. Oncogene. 2010;29:4914–4924. doi: 10.1038/onc.2010.237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 409.Ma YL, Zhang P, Wang F, Moyer MP, Yang JJ, Liu ZH, Peng JY, Chen HQ, Zhou YK, Liu WJ, Qin HL. Human embryonic stem cells and metastatic colorectal cancer cells shared the common endogenous human microRNA-26b. J Cell Mol Med. 2011;15:1941–1954. doi: 10.1111/j.1582-4934.2010.01170.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 410.Zhu Y, Lu Y, Zhang Q, Liu JJ, Li TJ, Yang JR, Zeng C, Zhuang SM. MicroRNA-26a/b and their host genes cooperate to inhibit the G1/S transition by activating the pRb protein. Nucleic Acids Res. doi: 10.1093/nar/gkr1278. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 411.Adlercreutz H, Honjo H, Higashi A, Fotsis T, Hamalainen E, Hasegawa T, Okada H. Urinary excretion of lignans and isoflavonoid phytoestrogens in Japanese men and women consuming a traditional Japanese diet. Am J Clin Nutr. 1991;54:1093–1100. doi: 10.1093/ajcn/54.6.1093. [DOI] [PubMed] [Google Scholar]
  • 412.Adlercreutz H, Markkanen H, Watanabe S. Plasma concentrations of phyto-oestrogens in Japanese men. Lancet. 1993;342:1209–1210. doi: 10.1016/0140-6736(93)92188-y. [DOI] [PubMed] [Google Scholar]
  • 413.Hebert JR, Hurley TG, Olendzki BC, Teas J, Ma Y, Hampl JS. Nutritional and socioeconomic factors in relation to prostate cancer mortality: a cross-national study. J Natl Cancer Inst. 1998;90:1637–1647. doi: 10.1093/jnci/90.21.1637. [DOI] [PubMed] [Google Scholar]
  • 414.Jacobsen BK, Knutsen SF, Fraser GE. Does high soy milk intake reduce prostate cancer incidence? The Adventist Health Study (United States) Cancer Causes Control. 1998;9:553–557. doi: 10.1023/a:1008819500080. [DOI] [PubMed] [Google Scholar]
  • 415.Kuang HB, Miao CL, Guo WX, Peng S, Cao YJ, Duan EK. Dickkopf-1 enhances migration of HEK293 cell by beta-catenin/E-cadherin degradation. Front Biosci. 2009;14:2212–2220. doi: 10.2741/3373. [DOI] [PubMed] [Google Scholar]
  • 416.Sarkar FH, Li Y. Harnessing the fruits of nature for the development of multi-targeted cancer therapeutics. Cancer Treat Rev. 2009;35:597–607. doi: 10.1016/j.ctrv.2009.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 417.Sarkar FH, Li Y, Wang Z, Padhye S. Lesson learned from nature for the development of novel anti-cancer agents: implication of isoflavone, curcumin, and their synthetic analogs. Curr Pharm Des. 2010;16:1801–1812. doi: 10.2174/138161210791208956. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 418.Su Y, Simmen FA, Xiao R, Simmen RC. Expression profiling of rat mammary epithelial cells reveals candidate signaling pathways in dietary protection from mammary tumors. Physiol Genomics. 2007;30:8–16. doi: 10.1152/physiolgenomics.00023.2007. [DOI] [PubMed] [Google Scholar]
  • 419.Su Y, Simmen RC. Soy isoflavone genistein upregulates epithelial adhesion molecule E-cadherin expression and attenuates beta-catenin signaling in mammary epithelial cells. Carcinogenesis. 2009;30:331–339. doi: 10.1093/carcin/bgn279. [DOI] [PubMed] [Google Scholar]
  • 420.Wang Z, Desmoulin S, Banerjee S, Kong D, Li Y, Deraniyagala RL, Abbruzzese J, Sarkar FH. Synergistic effects of multiple natural products in pancreatic cancer cells. Life Sci. 2008;83:293–300. doi: 10.1016/j.lfs.2008.06.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 421.Singh-Gupta V, Zhang H, Yunker CK, Ahmad Z, Zwier D, Sarkar FH, Hillman GG. Daidzein effect on hormone refractory prostate cancer in vitro and in vivo compared to genistein and soy extract: potentiation of radiotherapy. Pharm Res. 2010;27:1115–1127. doi: 10.1007/s11095-010-0107-9. [DOI] [PubMed] [Google Scholar]
  • 422.Singh-Gupta V, Zhang H, Banerjee S, Kong D, Raffoul JJ, Sarkar FH, Hillman GG. Radiation-induced HIF-1alpha cell survival pathway is inhibited by soy isoflavones in prostate cancer cells. Int J Cancer. 2009;124:1675–1684. doi: 10.1002/ijc.24015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 423.Bao B, Wang Z, Ali S, Kong D, Banerjee S, Ahmad A, Li Y, Azmi AS, Miele L, Sarkar FH. Over-expression of FoxM1 leads to epithelial–mesenchymal transition and cancer stem cell phenotype in pancreatic cancer cells. J Cell Biochem. 2011;112:2296–2306. doi: 10.1002/jcb.23150. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 424.Bao B, Wang Z, Ali S, Kong D, Li Y, Ahmad A, Banerjee S, Azmi AS, Miele L, Sarkar FH. Notch-1 induces epithelial–mesenchymal transition consistent with cancer stem cell phenotype in pancreatic cancer cells. Cancer Lett. 2011;307:26–36. doi: 10.1016/j.canlet.2011.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 425.Abas F, Lajis NH, Shaari K, Israf DA, Stanslas J, Yusuf UK, Raof SM. A labdane diterpene glucoside from the rhizomes of Curcuma mangga. J Nat Prod. 2005;68:1090–1093. doi: 10.1021/np0500171. [DOI] [PubMed] [Google Scholar]
  • 426.Hatcher H, Planalp R, Cho J, Torti FM, Torti SV. Curcumin: from ancient medicine to current clinical trials. Cell Mol Life Sci. 2008;65:1631–1652. doi: 10.1007/s00018-008-7452-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 427.Narayan S. Curcumin, a multi-functional chemopreventive agent, blocks growth of colon cancer cells by targeting beta-catenin-mediated transactivation and cell–cell adhesion pathways. J Mol Histol. 2004;35:301–307. doi: 10.1023/b:hijo.0000032361.98815.bb. [DOI] [PubMed] [Google Scholar]
  • 428.Mukherjee PK, Maity N, Nema NK, Sarkar BK. Bioactive compounds from natural resources against skin aging. Phytomedicine. 2011;19:64–73. doi: 10.1016/j.phymed.2011.10.003. [DOI] [PubMed] [Google Scholar]
  • 429.Padhye S, Chavan D, Pandey S, Deshpande J, Swamy KV, Sarkar FH. Perspectives on chemopreventive and therapeutic potential of curcumin analogs in medicinal chemistry. Mini Rev Med Chem. 2010;10:372–387. doi: 10.2174/138955710791330891. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 430.Wang Z, Li Y, Ahmad A, Banerjee S, Azmi AS, Kong D, Wojewoda C, Miele L, Sarkar FH. Down-regulation of Notch-1 is associated with Akt and FoxM1 in inducing cell growth inhibition and apoptosis in prostate cancer cells. J Cell Biochem. 2011;112:78–88. doi: 10.1002/jcb.22770. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 431.Upadhyay J, Kesharwani RK, Misra K. Comparative study of antioxidants as cancer preventives through inhibition of HIF-1 alpha activity. Bioinformation. 2009;4:233–236. doi: 10.6026/97320630004233. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 432.Bae MK, Kim SH, Jeong JW, Lee YM, Kim HS, Kim SR, Yun I, Bae SK, Kim KW. Curcumin inhibits hypoxia-induced angiogenesis via down-regulation of HIF-1. Oncol Rep. 2006;15:1557–1562. [PubMed] [Google Scholar]
  • 433.Strofer M, Jelkmann W, Depping R. Curcumin decreases survival of Hep3B liver and MCF-7 breast cancer cells: the role of HIF. Strahlenther Onkol. 2011;187:393–400. doi: 10.1007/s00066-011-2248-0. [DOI] [PubMed] [Google Scholar]
  • 434.Choi H, Chun YS, Kim SW, Kim MS, Park JW. Curcumin inhibits hypoxia-inducible factor-1 by degrading aryl hydrocarbon receptor nuclear translocator: a mechanism of tumor growth inhibition. Mol Pharmacol. 2006;70:1664–1671. doi: 10.1124/mol.106.025817. [DOI] [PubMed] [Google Scholar]
  • 435.Padhye S, Banerjee S, Chavan D, Pandye S, Swamy KV, Ali S, Li J, Dou QP, Sarkar FH. Fluorocurcumins as cyclooxygenase-2 inhibitor: molecular docking, pharmacokinetics and tissue distribution in mice. Pharm Res. 2009;26:2438–2445. doi: 10.1007/s11095-009-9955-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 436.Ali S, Ahmad A, Aboukameel A, Bao B, Padhye S, Philip PA, Sarkar FH. Increased Ras GTPase activity is regulated by miRNAs that can be attenuated by CDF treatment in pancreatic cancer cells. Cancer Lett. 2012;319:173–181. doi: 10.1016/j.canlet.2012.01.013. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 437.Mudduluru G, George-William JN, Muppala S, Asangani IA, Kumarswamy R, Nelson LD, Allgayer H. Curcumin regulates miR-21 expression and inhibits invasion and metastasis in colorectal cancer. Biosci Rep. 2011;31:185–197. doi: 10.1042/BSR20100065. [DOI] [PubMed] [Google Scholar]
  • 438.Reuter S, Gupta SC, Park B, Goel A, Aggarwal BB. Epigenetic changes induced by curcumin and other natural compounds. Genes Nutr. 2011;6:93–108. doi: 10.1007/s12263-011-0222-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 439.Jian L, Lee AH, Binns CW. Tea and lycopene protect against prostate cancer. Asia Pac J Clin Nutr. 2007;16(Suppl. 1):453–457. [PubMed] [Google Scholar]
  • 440.Kurahashi N, Sasazuki S, Iwasaki M, Inoue M, Tsugane S. Green tea consumption and prostate cancer risk in Japanese men: a prospective study. Am J Epidemiol. 2008;167:71–77. doi: 10.1093/aje/kwm249. [DOI] [PubMed] [Google Scholar]
  • 441.Khan N, Adhami VM, Mukhtar H. Review: green tea polyphenols in chemoprevention of prostate cancer: preclinical and clinical studies. Nutr Cancer. 2009;61:836–841. doi: 10.1080/01635580903285056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 442.Gao Z, Xu Z, Hung MS, Lin YC, Wang T, Gong M, Zhi X, Jablon DM, You L. Promoter demethylation of WIF-1 by epigallocatechin-3-gallate in lung cancer cells. Anticancer Res. 2009;29:2025–2030. [PubMed] [Google Scholar]
  • 443.Pahlke G, Ngiewih Y, Kern M, Jakobs S, Marko D, Eisenbrand G. Impact of quercetin and EGCG on key elements of the Wnt pathway in human colon carcinoma cells. J Agric Food Chem. 2006;54:7075–7082. doi: 10.1021/jf0612530. [DOI] [PubMed] [Google Scholar]
  • 444.Mount JG, Muzylak M, Allen S, Althnaian T, McGonnell IM, Price JS. Evidence that the canonical Wnt signalling pathway regulates deer antler regeneration. Dev Dyn. 2006;235:1390–1399. doi: 10.1002/dvdy.20742. [DOI] [PubMed] [Google Scholar]
  • 445.Bose M, Hao X, Ju J, Husain A, Park S, Lambert JD, Yang CS. Inhibition of tumorigenesis in ApcMin/+ mice by a combination of (−)-epigallocatechin-3-gallate and fish oil. J Agric Food Chem. 2007;55:7695–7700. doi: 10.1021/jf071004r. [DOI] [PubMed] [Google Scholar]
  • 446.Liu L, Lai CQ, Nie L, Ordovas J, Band M, Moser L, Meydani M. The modulation of endothelial cell gene expression by green tea polyphenol-EGCG. Mol Nutr Food Res. 2008;52:1182–1192. doi: 10.1002/mnfr.200700499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 447.Kim J, Zhang X, Rieger-Christ KM, Summerhayes IC, Wazer DE, Paulson KE, Yee AS. Suppression of Wnt signaling by the green tea compound (–)-epigallocatechin 3-gallate (EGCG) in invasive breast cancer cells. Requirement of the transcriptional repressor HBP1. J Biol Chem. 2006;281:10865–10875. doi: 10.1074/jbc.M513378200. [DOI] [PubMed] [Google Scholar]
  • 448.Tang GQ, Yan TQ, Guo W, Ren TT, Peng CL, Zhao H, Lu XC, Zhao FL, Han X. (–)-Epigallocatechin-3-gallate induces apoptosis and suppresses proliferation by inhibiting the human Indian Hedgehog pathway in human chondrosarcoma cells. J Cancer Res Clin Oncol. 2010;136:1179–1185. doi: 10.1007/s00432-010-0765-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 449.Domingo DS, Camouse MM, Hsia AH, Matsui M, Maes D, Ward NL, Cooper KD, Baron ED. Anti-angiogenic effects of epigallocatechin-3-gallate in human skin. Int J Clin Exp Pathol. 2010;3:705–709. [PMC free article] [PubMed] [Google Scholar]
  • 450.Shimizu M, Shirakami Y, Sakai H, Yasuda Y, Kubota M, Adachi S, Tsurumi H, Hara Y, Moriwaki H. (−)-Epigallocatechin gallate inhibits growth and activation of the VEGF/VEGFR axis in human colorectal cancer cells. Chem Biol Interact. 2010;185:247–252. doi: 10.1016/j.cbi.2010.03.036. [DOI] [PubMed] [Google Scholar]
  • 451.Lin SK, Shun CT, Kok SH, Wang CC, Hsiao TY, Liu CM. Hypoxia-stimulated vascular endothelial growth factor production in human nasal polyp fibroblasts: effect of epigallocatechin-3-gallate on hypoxia-inducible factor-1 alpha synthesis. Arch Otolaryngol Head Neck Surg. 2008;134:522–527. doi: 10.1001/archotol.134.5.522. [DOI] [PubMed] [Google Scholar]
  • 452.Zhang Q, Tang X, Lu Q, Zhang Z, Rao J, Le AD. Green tea extract and (−)-epigallocatechin-3-gallate inhibit hypoxia- and serum-induced HIF-1alpha protein accumulation and VEGF expression in human cervical carcinoma and hepatoma cells. Mol Cancer Ther. 2006;5:1227–1238. doi: 10.1158/1535-7163.MCT-05-0490. [DOI] [PubMed] [Google Scholar]
  • 453.Tsang WP, Kwok TT. Epigallocatechin gallate up-regulation of miR-16 and induction of apoptosis in human cancer cells. J Nutr Biochem. 2010;21:140–146. doi: 10.1016/j.jnutbio.2008.12.003. [DOI] [PubMed] [Google Scholar]
  • 454.Manalo DJ, Baek JH, Buehler PW, Struble E, Abraham B, Alayash AI. Inactivation of prolyl hydroxylase domain (PHD) protein by epigallocatechin (EGCG) stabilizes hypoxia-inducible factor (HIF-1alpha) and induces hepcidin (Hamp) in rat kidney. Biochem Biophys Res Commun. 2011;416:421–426. doi: 10.1016/j.bbrc.2011.11.085. [DOI] [PubMed] [Google Scholar]
  • 455.Wang H, Bian S, Yang CS. Green tea polyphenol EGCG suppresses lung cancer cell growth through upregulating miR-210 expression caused by stabilizing HIF-1alpha. Carcinogenesis. 2011;32:1881–1889. doi: 10.1093/carcin/bgr218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 456.Thomas R, Kim MH. Epigallocatechin gallate inhibits HIF-1alpha degradation in prostate cancer cells. Biochem Biophys Res Commun. 2005;334:543–548. doi: 10.1016/j.bbrc.2005.06.114. [DOI] [PubMed] [Google Scholar]
  • 457.Zhou YD, Kim YP, Li XC, Baerson SR, Agarwal AK, Hodges TW, Ferreira D, Nagle DG. Hypoxia-inducible factor-1 activation by (−)-epicatechin gallate: potential adverse effects of cancer chemoprevention with high-dose green tea extracts. J Nat Prod. 2004;67:2063–2069. doi: 10.1021/np040140c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 458.Shakibaei M, Harikumar KB, Aggarwal BB. Resveratrol addiction: to die or not to die. Mol Nutr Food Res. 2009;53:115–128. doi: 10.1002/mnfr.200800148. [DOI] [PubMed] [Google Scholar]
  • 459.Vanamala J, Reddivari L, Radhakrishnan S, Tarver C. Resveratrol suppresses IGF-1 induced human colon cancer cell proliferation and elevates apoptosis via suppression of IGF-1R/Wnt and activation of p53 signaling pathways. BMC Cancer. 2010;10:238. doi: 10.1186/1471-2407-10-238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 460.Hope C, Planutis K, Planutiene M, Moyer MP, Johal KS, Woo J, Santoso C, Hanson JA, Holcombe RF. Low concentrations of resveratrol inhibit Wnt signal throughput in colon-derived cells: implications for colon cancer prevention. Mol Nutr Food Res. 2008;52(Suppl. 1):S52–S61. doi: 10.1002/mnfr.200700448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 461.Roccaro AM, Leleu X, Sacco A, Moreau AS, Hatjiharissi E, Jia X, Xu L, Ciccarelli B, Patterson CJ, Ngo HT, Russo D, Vacca A, Dammacco F, Anderson KC, Ghobrial IM, Treon SP. Resveratrol exerts antiproliferative activity and induces apoptosis in Waldenstrom’s macroglobulinemia. Clin Cancer Res. 2008;14:1849–1858. doi: 10.1158/1078-0432.CCR-07-1750. [DOI] [PubMed] [Google Scholar]
  • 462.Cho SW, Her SJ, Sun HJ, Choi OK, Yang JY, Kim SW, Kim SY, Shin CS. Differential effects of secreted frizzled-related proteins (sFRPs) on osteoblastic differentiation of mouse mesenchymal cells and apoptosis of osteoblasts. Biochem Biophys Res Commun. 2008;367:399–405. doi: 10.1016/j.bbrc.2007.12.128. [DOI] [PubMed] [Google Scholar]
  • 463.Cho SW, Yang JY, Sun HJ, Jung JY, Her SJ, Cho HY, Choi HJ, Kim SW, Kim SY, Shin CS. Wnt inhibitory factor (WIF)-1 inhibits osteoblastic differentiation in mouse embryonic mesenchymal cells. Bone. 2009;44:1069–1077. doi: 10.1016/j.bone.2009.02.012. [DOI] [PubMed] [Google Scholar]
  • 464.Zhou H, Shang L, Li X, Zhang X, Gao G, Guo C, Chen B, Liu Q, Gong Y, Shao C. Resveratrol augments the canonical Wnt signaling pathway in promoting osteoblastic differentiation of multipotent mesenchymal cells. Exp Cell Res. 2009;315:2953–2962. doi: 10.1016/j.yexcr.2009.07.030. [DOI] [PubMed] [Google Scholar]
  • 465.Wang S, Qian Y, Gong D, Zhang Y, Fan Y. Resveratrol attenuates acute hypoxic injury in cardiomyocytes: correlation with inhibition of iNOS-NO signaling pathway. Eur J Pharm Sci. 2011;44:416–421. doi: 10.1016/j.ejps.2011.08.029. [DOI] [PubMed] [Google Scholar]
  • 466.Agrawal M, Kumar V, Kashyap MP, Khanna VK, Randhawa GS, Pant AB. Ischemic insult induced apoptotic changes in PC12 cells: protection by trans resveratrol. Eur J Pharmacol. 2011;666:5–11. doi: 10.1016/j.ejphar.2011.05.015. [DOI] [PubMed] [Google Scholar]
  • 467.Huang X, Zhou Q, Gu L, Liu D, Li Z, Liu Q, Zhu D. Protective effects of resveratrol on neonatal rat cardiomyocyte lesion induced by hypoxia. Zhongguo Zhong Yao Za Zhi. 2010;35:94–98. [PubMed] [Google Scholar]
  • 468.Quan F, Zhang SQ, Bai YX, Yao XB, Li HH, Yu L, Pan CE. Resveratrol increases sensitivity of CNE2 cells to chemotherapeutic drugs under hypoxia. Zhong Xi Yi Jie He Xue Bao. 2009;7:952–957. doi: 10.3736/jcim20091008. [DOI] [PubMed] [Google Scholar]
  • 469.Wu H, Liang X, Fang Y, Qin X, Zhang Y, Liu J. Resveratrol inhibits hypoxia-induced metastasis potential enhancement by restricting hypoxia-induced factor-1 alpha expression in colon carcinoma cells. Biomed Pharmacother. 2008;62:613–621. doi: 10.1016/j.biopha.2008.06.036. [DOI] [PubMed] [Google Scholar]
  • 470.Park SY, Jeong KJ, Lee J, Yoon DS, Choi WS, Kim YK, Han JW, Kim YM, Kim BK, Lee HY. Hypoxia enhances LPA-induced HIF-1alpha and VEGF expression: their inhibition by resveratrol. Cancer Lett. 2007;258:63–69. doi: 10.1016/j.canlet.2007.08.011. [DOI] [PubMed] [Google Scholar]
  • 471.Bae S, Lee EM, Cha HJ, Kim K, Yoon Y, Lee H, Kim J, Kim YJ, Lee HG, Jeung HK, Min YH, An S. Resveratrol alters microRNA expression profiles in A549 human non-small cell lung cancer cells. Mol Cells. 2011;32:243–249. doi: 10.1007/s10059-011-1037-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 472.Tili E, Michaille JJ, Alder H, Volinia S, Delmas D, Latruffe N, Croce CM. Resveratrol modulates the levels of microRNAs targeting genes encoding tumor-suppressors and effectors of TGFbeta signaling pathway in SW480 cells. Biochem Pharmacol. 2010;80:2057–2065. doi: 10.1016/j.bcp.2010.07.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 473.Higdon JV, Delage B, Williams DE, Dashwood RH. Cruciferous vegetables and human cancer risk: epidemiologic evidence and mechanistic basis. Pharmacol Res. 2007;55:224–236. doi: 10.1016/j.phrs.2007.01.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 474.Nho CW, Jeffery E. Crambene, a bioactive nitrile derived from glucosinolate hydrolysis, acts via the antioxidant response element to upregulate quinone reductase alone or synergistically with indole-3-carbinol. Toxicol Appl Pharmacol. 2004;198:40–48. doi: 10.1016/j.taap.2004.02.012. [DOI] [PubMed] [Google Scholar]
  • 475.Benabadji SH, Wen R, Zheng JB, Dong XC, Yuan SG. Anticarcinogenic and antioxidant activity of diindolylmethane derivatives. Acta Pharmacol Sin. 2004;25:666–671. [PubMed] [Google Scholar]
  • 476.Fares F, Azzam N, Appel B, Fares B, Stein A. The potential efficacy of 3,3′-diindolylmethane in prevention of prostate cancer development. Eur J Cancer Prev. 2010;19:199–203. doi: 10.1097/CEJ.0b013e328333fbce. [DOI] [PubMed] [Google Scholar]
  • 477.Li Y, Wang Z, Kong D, Murthy S, Dou QP, Sheng S, Reddy GP, Sarkar FH. Regulation of FOXO3a/beta-catenin/GSK-3beta signaling by 3,3′-diindolylmethane contributes to inhibition of cell proliferation and induction of apoptosis in prostate cancer cells. J Biol Chem. 2007;282:21542–21550. doi: 10.1074/jbc.M701978200. [DOI] [PubMed] [Google Scholar]
  • 478.Li Y, Zhang T, Korkaya H, Liu S, Lee HF, Newman B, Yu Y, Clouthier SG, Schwartz SJ, Wicha MS, Sun D. Sulforaphane, a dietary component of broccoli/broccoli sprouts, inhibits breast cancer stem cells. Clin Cancer Res. 2010;16:2580–2590. doi: 10.1158/1078-0432.CCR-09-2937. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 479.Riby JE, Firestone GL, Bjeldanes LF. 3,3′-Diindolylmethane reduces levels of HIF-1alpha and HIF-1 activity in hypoxic cultured human cancer cells. Biochem Pharmacol. 2008;75:1858–1867. doi: 10.1016/j.bcp.2008.01.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 480.Singh-Gupta V, Banerjee S, Yunker CK, Rakowski JT, Joiner MC, Konski AA, Sarkar FH, Hillman GG. B-DIM impairs radiation-induced survival pathways independently of androgen receptor expression and augments radiation efficacy in prostate cancer. Cancer Lett. 2012;318:86–92. doi: 10.1016/j.canlet.2011.12.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 481.Jin Y, Zou X, Feng X. 3,3′-Diindolylmethane negatively regulates Cdc25A and induces a G2/M arrest by modulation of microRNA 21 in human breast cancer cells. Anticancer Drugs. 2010;21:814–822. doi: 10.1097/CAD.0b013e32833e53ea. [DOI] [PubMed] [Google Scholar]
  • 482.Giovannucci E, Rimm EB, Liu Y, Stampfer MJ, Willett WC. A prospective study of tomato products, lycopene, and prostate cancer risk. J Natl Cancer Inst. 2002;94:391–398. doi: 10.1093/jnci/94.5.391. [DOI] [PubMed] [Google Scholar]
  • 483.Gann PH, Ma J, Giovannucci E, Willett W, Sacks FM, Hennekens CH, Stampfer MJ. Lower prostate cancer risk in men with elevated plasma lycopene levels: results of a prospective analysis. Cancer Res. 1999;59:1225–1230. [PubMed] [Google Scholar]
  • 484.Lu QY, Hung JC, Heber D, Go VL, Reuter VE, Cordon-Cardo C, Scher HI, Marshall JR, Zhang ZF. Inverse associations between plasma lycopene and other carotenoids and prostate cancer. Cancer Epidemiol Biomarkers Prev. 2001;10:749–756. [PubMed] [Google Scholar]
  • 485.Limpens J, van Weerden WM, Kramer K, Pallapies D, Obermuller-Jevic UC, Schroder FH. Re: Prostate carcinogenesis in N-methyl-N-nitrosourea (NMU)-testosterone-treated rats fed tomato powder, lycopene, or energy-restricted diets. J Natl Cancer Inst. 2004;96:554–555. doi: 10.1093/jnci/djh089. [DOI] [PubMed] [Google Scholar]
  • 486.Kucuk O, Sarkar FH, Sakr W, Djuric Z, Pollak MN, Khachik F, Li YW, Banerjee M, Grignon D, Bertram JS, Crissman JD, Pontes EJ, Wood DP., Jr Phase II randomized clinical trial of lycopene supplementation before radical prostatectomy. Cancer Epidemiol Biomarkers Prev. 2001;10:861–868. [PubMed] [Google Scholar]
  • 487.Kucuk O, Sarkar FH, Djuric Z, Sakr W, Pollak MN, Khachik F, Banerjee M, Bertram JS, Wood DP., Jr Effects of lycopene supplementation in patients with localized prostate cancer. Exp Biol Med (Maywood) 2002;227:881–885. doi: 10.1177/153537020222701007. [DOI] [PubMed] [Google Scholar]
  • 488.Wertz K. Lycopene effects contributing to prostate health. Nutr Cancer. 2009;61:775–783. doi: 10.1080/01635580903285023. [DOI] [PubMed] [Google Scholar]
  • 489.Wei Y, Shen X, Shen H, Mai J, Wu M, Yao G. Effects of lycopene on cerebral ischemia-reperfusion injury in rats. Wei Sheng Yan Jiu. 2010;39:201–204. [PubMed] [Google Scholar]
  • 490.Garland CF, Garland FC, Gorham ED, Lipkin M, Newmark H, Mohr SB, Holick MF. The role of vitamin D in cancer prevention. Am J Public Health. 2006;96:252–261. doi: 10.2105/AJPH.2004.045260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 491.Ahn J, Albanes D, Peters U, Schatzkin A, Lim U, Freedman M, Chatterjee N, Andriole GL, Leitzmann MF, Hayes RB. Dairy products, calcium intake, and risk of prostate cancer in the prostate, lung, colorectal, and ovarian cancer screening trial. Cancer Epidemiol Biomarkers Prev. 2007;16:2623–2630. doi: 10.1158/1055-9965.EPI-07-0601. [DOI] [PubMed] [Google Scholar]
  • 492.Ben-Shoshan M, Amir S, Dang DT, Dang LH, Weisman Y, Mabjeesh NJ. 1Alpha,25-dihydroxyvitamin D3 (Calcitriol) inhibits hypoxia-inducible factor-1/vascular endothelial growth factor pathway in human cancer cells. Mol Cancer Ther. 2007;6:1433–1439. doi: 10.1158/1535-7163.MCT-06-0677. [DOI] [PubMed] [Google Scholar]
  • 493.Pike JW, Meyer MB, Martowicz ML, Bishop KA, Lee SM, Nerenz RD, Goetsch PD. Emerging regulatory paradigms for control of gene expression by 1,25-dihydroxyvitamin D3. J Steroid Biochem Mol Biol. 2010;121:130–135. doi: 10.1016/j.jsbmb.2010.02.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 494.Jiang Y, Zheng W, Teegarden D. 1Alpha,25-dihydroxyvitamin D regulates hypoxia-inducible factor-1alpha in untransformed and Harvey-ras transfected breast epithelial cells. Cancer Lett. 2010;298:159–166. doi: 10.1016/j.canlet.2010.06.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 495.Wang WL, Chatterjee N, Chittur SV, Welsh J, Tenniswood MP. Effects of 1alpha, 25 dihydroxyvitamin D3 and testosterone on miRNA and mRNA expression in LNCaP cells. Mol Cancer. 2011;10:58. doi: 10.1186/1476-4598-10-58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 496.Lutherborrow M, Bryant A, Jayaswal V, Agapiou D, Palma C, Yang YH, Ma DD. Expression profiling of cytogenetically normal acute myeloid leukemia identifies microRNAs that target genes involved in monocytic differentiation. Am J Hematol. 2011;86:2–11. doi: 10.1002/ajh.21864. [DOI] [PubMed] [Google Scholar]
  • 497.Isham CR, Tibodeau JD, Jin W, Xu R, Timm MM, Bible KC. Chaetocin: a promising new antimyeloma agent with in vitro and in vivo activity mediated via imposition of oxidative stress. Blood. 2007;109:2579–2588. doi: 10.1182/blood-2006-07-027326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 498.Ban HS, Uto Y, Nakamura H. Hypoxia-inducible factor inhibitors: a survey of recent patented compounds (2004–2010) Expert Opin Ther Pat. 2011;21:131–146. doi: 10.1517/13543776.2011.547477. [DOI] [PubMed] [Google Scholar]
  • 499.Lee YM, Lim JH, Yoon H, Chun YS, Park JW. Antihepatoma activity of chaetocin due to deregulated splicing of hypoxia-inducible factor 1alpha pre-mRNA in mice and in vitro. Hepatology. 2011;53:171–180. doi: 10.1002/hep.24010. [DOI] [PubMed] [Google Scholar]
  • 500.Rao KV, Puri VN, Diwan PK, Alvarez FM. Preliminary evaluation of manassantin A, a potential neuroleptic agent from Saururus cernuus. Pharmacol Res Commun. 1987;19:629–638. doi: 10.1016/0031-6989(87)90117-2. [DOI] [PubMed] [Google Scholar]
  • 501.Hwang BY, Lee JH, Nam JB, Hong YS, Lee JJ. Lignans from Saururus chinensis inhibiting the transcription factor NF-kappaB. Phytochemistry. 2003;64:765–771. doi: 10.1016/s0031-9422(03)00391-1. [DOI] [PubMed] [Google Scholar]
  • 502.Lee WS, Lee DW, Baek YI, An S, Cho KH, Choi YK, Kim HC, Park HY, Bae KH, Jeong TS. Human ACAT-1 and -2 inhibitory activities of saucerneol B, manassantin A and B isolated from Saururus chinensis. Bioorg Med Chem Lett. 2004;14:3109–3112. doi: 10.1016/j.bmcl.2004.04.023. [DOI] [PubMed] [Google Scholar]
  • 503.Lee JH, Hwang BY, Kim KS, Nam JB, Hong YS, Lee JJ. Suppression of RelA/p65 transactivation activity by a lignoid manassantin isolated from Saururus chinensis. Biochem Pharmacol. 2003;66:1925–1933. doi: 10.1016/s0006-2952(03)00553-7. [DOI] [PubMed] [Google Scholar]
  • 504.Kasper AC, Moon EJ, Hu X, Park Y, Wooten CM, Kim H, Yang W, Dewhirst MW, Hong J. Analysis of HIF-1 inhibition bymanassantinA and analogues withmodified tetrahydrofuran configurations. Bioorg Med Chem Lett. 2009;19:3783–3786. doi: 10.1016/j.bmcl.2009.04.071. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 505.Hodges TW, Hossain CF, Kim YP, Zhou YD, Nagle DG. Molecular-targeted antitumor agents: the Saururus cernuus dineolignans manassantin B and 4-O-demethylmanassantin B are potent inhibitors of hypoxia-activated HIF-1. J Nat Prod. 2004;67:767–771. doi: 10.1021/np030514m. [DOI] [PubMed] [Google Scholar]
  • 506.Fraker PJ, Gershwin ME, Good RA, Prasad A. Interrelationships between zinc and immune function. Fed Proc. 1986;45:1474–1479. [PubMed] [Google Scholar]
  • 507.Fraker PJ, Telford WG. A reappraisal of the role of zinc in life and death decisions of cells. Proc Soc Exp Biol Med. 1997;215:229–236. doi: 10.3181/00379727-215-44132. [DOI] [PubMed] [Google Scholar]
  • 508.Allen JI, Bell E, Boosalis MG, Oken MM, McClain CJ, Levine AS, Morley JE. Association between urinary zinc excretion and lymphocyte dysfunction in patients with lung cancer. Am J Med. 1985;79:209–215. doi: 10.1016/0002-9343(85)90011-7. [DOI] [PubMed] [Google Scholar]
  • 509.Feustel A, Wennrich R, Steiniger D, Klauss P. Zinc and cadmium concentration in prostatic carcinoma of different histological grading in comparison to normal prostate tissue and adenofibromyomatosis (BPH) Urol Res. 1982;10:301–303. doi: 10.1007/BF00255877. [DOI] [PubMed] [Google Scholar]
  • 510.Feustel A, Wennrich R, Schmidt B. Serum-Zn-levels in prostatic cancer. Urol Res. 1989;17:41–42. doi: 10.1007/BF00261049. [DOI] [PubMed] [Google Scholar]
  • 511.Ogunlewe JO, Osegbe DN. Zinc and cadmium concentrations in indigenous blacks with normal, hypertrophic, and malignant prostate. Cancer. 1989;63:1388–1392. doi: 10.1002/1097-0142(19890401)63:7<1388::aid-cncr2820630725>3.0.co;2-m. [DOI] [PubMed] [Google Scholar]
  • 512.Yang CS. Research on esophageal cancer in China: a review. Cancer Res. 1980;40:2633–2644. [PubMed] [Google Scholar]
  • 513.Prasad AS. Zinc: an overview. Nutrition. 1995;11:93–99. [PubMed] [Google Scholar]
  • 514.Prasad AS, Kucuk O. Zinc in cancer prevention. Cancer Metastasis Rev. 2002;21:291–295. doi: 10.1023/a:1021215111729. [DOI] [PubMed] [Google Scholar]
  • 515.Prasad AS, Mukhtar H, Beck FW, Adhami VM, Siddiqui IA, Din M, Hafeez BB, Kucuk O. Dietary zinc and prostate cancer in the TRAMP mouse model. J Med Food. 2010;13:70–76. doi: 10.1089/jmf.2009.0042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 516.Prasad AS, Bao B, Beck FW, Sarkar FH. Zinc-suppressed inflammatory cytokines by induction of A20-mediated inhibition of nuclear factor-kappaB. Nutrition. 2011;27:816–823. doi: 10.1016/j.nut.2010.08.010. [DOI] [PubMed] [Google Scholar]
  • 517.Kim I, Kim CH, Seo GH, Kim HS, Lee J, Kim DG, Ahn YS. Inhibitory effect of zinc on hypoxic HIF-1 activation in astrocytes. Neuroreport. 2008;19:1063–1066. doi: 10.1097/WNR.0b013e328304d9ac. [DOI] [PubMed] [Google Scholar]

RESOURCES