Next Article in Journal
Cost-Effectiveness Analysis of Axicabtagene Ciloleucel vs. Tisagenlecleucel for the Management of Relapsed/Refractory Diffuse Large B-Cell Lymphoma in Spain
Previous Article in Journal
Alternative NF-κB Signaling Discriminates Induction of the Tumor Marker Fascin by the Viral Oncoproteins Tax-1 and Tax-2 of Human T-Cell Leukemia Viruses
Previous Article in Special Issue
Role of FGFR2c and Its PKCε Downstream Signaling in the Control of EMT and Autophagy in Pancreatic Ductal Adenocarcinoma Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Multifaceted Interplay between Hormones, Growth Factors and Hypoxia in the Tumor Microenvironment

1
Department of Pharmacy, Health and Nutritional Sciences, University of Calabria, 87036 Rende, Italy
2
Department of Biology, University of Waterloo, Waterloo, ON N2L 3G1, Canada
3
Department of Laboratory Medicine & Pathobiology, University of Toronto, Toronto, ON M5S 1A8, Canada
4
Kite Pharma Inc., Santa Monica, CA 90404, USA
*
Authors to whom correspondence should be addressed.
Cancers 2022, 14(3), 539; https://doi.org/10.3390/cancers14030539
Submission received: 23 December 2021 / Revised: 17 January 2022 / Accepted: 18 January 2022 / Published: 21 January 2022

Abstract

:

Simple Summary

Hormones and growth factors impact many processes in the cell. Moreover, these molecules influence tumor growth, as does a lack of oxygen (hypoxia) that characterizes cancer progression. Proteins that are stabilized by low oxygen tension, known as hypoxia-inducible factors (HIFs), help tumor cells to adapt to their environment. Of note, hormones and growth factors regulate the activity of HIFs toward malignant aggressiveness, including the resistance to therapy. In this review, we summarize the current knowledge regarding the role of hormones and growth factors in cancer development with a particular focus on their interplay with hypoxia and HIFs and comment on how these factors influence the response to cancer immunotherapy.

Abstract

Hormones and growth factors (GFs) are signaling molecules implicated in the regulation of a variety of cellular processes. They play important roles in both healthy and tumor cells, where they function by binding to specific receptors on target cells and activating downstream signaling cascades. The stages of tumor progression are influenced by hormones and GF signaling. Hypoxia, a hallmark of cancer progression, contributes to tumor plasticity and heterogeneity. Most solid tumors contain a hypoxic core due to rapid cellular proliferation that outgrows the blood supply. In these circumstances, hypoxia-inducible factors (HIFs) play a central role in the adaptation of tumor cells to their new environment, dramatically reshaping their transcriptional profile. HIF signaling is modulated by a variety of factors including hormones and GFs, which activate signaling pathways that enhance tumor growth and metastatic potential and impair responses to therapy. In this review, we summarize the role of hormones and GFs during cancer onset and progression with a particular focus on hypoxia and the interplay with HIF proteins. We also discuss how hypoxia influences the efficacy of cancer immunotherapy, considering that a hypoxic environment may act as a determinant of the immune-excluded phenotype and a major hindrance to the success of adoptive cell therapies.

1. Importance of Hormones and Growth Factors (GFs) in Tumor Onset and Progression

Hormones and GFs are described as endogenous chemical components that act as cellular signaling factors to regulate a variety of processes such as cell growth, maturation, differentiation, function, and metabolism [1,2]. They exert their functions by binding to specific receptors on target cells to activate downstream signaling cascades. This activation can suppress or enhance the aforementioned processes. Such cellular responses are frequently described as negative or positive feedback loops [2].
Hormones and GFs are essential to the regulation of crucial processes in healthy cells. However, they also significantly affect metabolic processes in tumor cells. Moreover, despite having a direct effect on single tumor cells, hormones and GFs can also mediate the interplay between tumor cells, their interaction with the extracellular matrix (ECM), as well as their interaction with cells of surrounding tissues [3]. Such complex interactions significantly affect processes that influence tumor behavior such as proliferation, angiogenesis, or local inflammatory responses [4,5].
Many tumor cells frequently do not depend on the synthesis of signaling molecules from neighboring cells but can synthetize them independently and thus mediate their own metabolism and proliferation [6]. The secretion of hormones and GFs by cancer cells is also known as ectopic secretion and is often associated with tumors of endocrine tissues such as pancreatic islet cells, parathyroid cells, thyroid cells, or neuroendocrine cells that can be found in nearly every organ [7,8]. Hormones and GFs that influence tumor growth and progression can be classified into several groups, amongst the most important being steroid hormones such as estrogens and androgens, Epidermal GFs (EGFs) and Neuregulins (NRGs), Insulin-like GFs (IGFs), Transforming GF-β (TGF-β), and Vascular Endothelial GFs (VEGFs) [5].
Thus, hormones and GFs can affect various aspects of tumor cell metabolism. However, considering that tumors can arise from various cell and tissue types, different tumors can activate different molecular players and mechanisms to facilitate their growth and progression. The following sections will provide an overview of the key hormones and GFs, pathways, and mechanisms driving tumor cell metabolism.

1.1. Hormone and GF-Mediated Regulation of Intracellular Signaling Cascades

Hormones that affect the growth of tumor cells are usually lipid-soluble molecules that can pass directly through the cellular membrane and bind to intracellular proteins or nuclear receptors to mediate downstream signaling events and gene expression. In contrast, most GFs require engagement with a specific cell-surface receptor to enter the cell and mediate cellular signaling [9]. Accordingly, hormones mainly influence the expression of target genes, whereas GFs can directly mediate processes that take place in the cytoplasm. Moreover, GFs can also influence gene expression. Despite activating classical signaling cascades, GFs can indirectly or directly function as transcription factors, as well as induce the activity of transcription factors, like Signal Transducers and Activators of Transcription (STATs), or SMADs, and thus mediate gene expression [5,10,11,12,13]. Thus, both hormones and GFs ultimately mediate important cellular processes [14]. For example, upon GF-receptor engagement or the activation of hormone-responsive genes, protein phosphorylation events and protein kinase activities are crucial in the transmission of growth signals [9,14]. Prominent signaling cascades that are subsequently activated include various mitogenic pathways such as the MAPK/ERK, SMAD, PI3K/AKT, Ras-like GTPases, or the phospholipase C-γ pathways [9].
Overall, ligand-receptor binding is a crucial step in the activation of molecular signaling cascades. A prominent example of ligand-receptor binding that influences cellular growth is the activation of the Growth Hormone Receptor (GHR) upon Growth Hormone (GH) binding. Their engagement results in phosphorylation of the receptor-associated kinase Janus Kinase-2 (JAK2) and activation of STAT5, which are known mediators of GH function and oncogenesis [13,15]. GH has also been demonstrated to affect the tissue-specific expression of microRNAs, indicating a novel role for GH beyond its classical function in the activation of signaling pathways [13,16]. Another example of GF binding that affects downstream cellular signaling pathways that underlie the development of prominent tumor signatures is the binding of IGF-I to its receptor in the PI3K/AKT pathway. Once activated, PI3K phosphorylates and activates the serine/threonine-specific protein kinase AKT. AKT activation mediates processes that increase cellular growth, cell survival, cell motility, and angiogenesis, as well as the inhibition of apoptosis. Accordingly, many tumor cells exhibit a resistance to apoptosis that lends them a significant survival advantage compared to non-tumor cells [9]. Moreover, angiogenesis is important to the growth of many tumors. For instance, important GFs that affect angiogenesis include the VEGF family and EGF. Blocking them from binding to tumor cells can cause tumor-cell death [9,17,18,19]. Thus, interfering with the binding of hormones and GFs to tumor cells may represent an important strategy for combating tumor development and progression, respectively.
Taken together, the binding of hormones and GFs can have direct effects on metabolic processes in target cells. Considering that most tumors develop from the clonal expansion of one cell mutated via somatic mutation [5,20], a process that is strongly influenced by the binding of signaling molecules, it is clear that key steps of tumor development and progression are driven by hormones and GFs [5,9,15]. The following section will describe how hormones and GFs influence the stages of tumor growth and progression.

1.2. Hormones and GFs in Tumor Growth and Progression

Tumor growth and progression represent a cascade of different developmental stages that are influenced by the interplay between a variety of factors in the tumor’s intra- and extracellular environment. The process of tumor progression can be divided into several distinct steps: (1) somatic mutation, (2) clonal expansion, (3) intraluminal cell proliferation and intraluminal lesion formation, (4) invasion, (5) dissemination, (6) the formation of micrometastases, (7) resistant tumor clones, (8) angiogenesis, and ultimately (9) metastases [5] (Figure 1). These stages of tumor growth and progression are substantially affected by the hormone and GF signaling [5,9,13,15].
Upon the initial somatic mutation and transformation of a healthy cell into a tumor cell, processes of cell growth and expansion are initiated. Tumor cells can mediate their own growth by the ectopic secretion of hormones and GFs [6]. In addition, despite their ability to perform ectopic secretion, tumor cells are frequently characterized by an overexpression of GF receptors, which contributes to their increased responsiveness to GFs. For instance, the EGF Receptor (EGFR) is commonly overexpressed in head or neck cancer cells, and ErbB-2/HER2 is overexpressed in breast cancer cells [5,21]. Moreover, signaling factors involved in various mitogenic pathways can be mutated in different tumor cells, which typically increase tumor growth. For example, RAS is mutated in 25% of human cancers [5].
Somatic mutation is followed by the clonal expansion of the tumor cell, which is frequently supported by GFs such as EGF and IGF-I [22]. Clonal expansion often contributes to intraluminal lesions in the organ [23]. Intraluminal cell proliferation and the formation of intraluminal lesions are frequently associated with abnormal hormonal and GF signaling. For example, ErbB-2/HER2 is commonly overexpressed in 40% of ductal carcinoma in situ (DCIS) breast cancer cases [24]. Intraluminal lesions occur upon the disruption of the basal cellular membrane, which represents the basis for tumor progression. The formation of intraluminal lesions is followed by the invasion of neighboring tissues and organs by tumor cells [25,26]. Invasion is possible due to acquired increased motility, loss of epithelial polarity, protease secretion, and the epithelial-mesenchymal transition (EMT) of tumor cells. Features of tumor-cell invasiveness are controlled by a multitude of hormones and GFs. For example, the activation of the Estrogen Receptors (ERs) by estrogens and Progesterone Receptors (PRs) by progesterone in transformed cells directly affects the cellular proliferation status, intracellular adhesion, cell motility, and morphology [14] via the activation of multiple transcription factors that alter the expression of genes involved in these processes. Moreover, signaling factors like the TGF-β or autocrine stimulatory loop signaling via EGFR have been demonstrated to promote tumor invasiveness [27,28] by causing a shift in cell adhesion properties of tumor cells, mainly through the downregulation of protease inhibitors, upregulation of protease secretion, or by causing a general increase in cell motility [5,28,29,30].
Tumor invasion is enhanced due to the EMT of tumor cells. The EMT is characterized by a mesenchymal stem cell-like stage of tumor cells, where initially polarized cells are transformed into non-polarized cells with increased motility that secrete ECM components [5,31]. A hallmark of cells that have undergone EMT is the loss of E-cadherin due to its mutational inactivation, transcriptional repression, or proteolysis [5]. E-cadherin is a key regulator of cell adhesion and the maintenance of an epithelial-cell phenotype [32]. In tumor cells, GFs like TGF-β, EGF, or Notch proteins regulate the expression of E-cadherin repressors [5,32,33] to maintain their stem cell-like phenotype.
Following invasion, tumor dissemination can take place. Dissemination is described as the process of tumor-cells spreading into other regions, or organs of the body via the entry (extravasation) and exit (intravasation) through the lymphatic, capillary, or blood vessel system [5,34,35]. Tumor dissemination is significantly influenced by signaling between tumor cells and surrounding cells such as endothelial cells and macrophages. For example, paracrine signaling between different cell types in the tumor microenvironment (TME), as well as the proinflammatory and hypoxic conditions during tumor progression, can facilitate the increased secretion of TGF-β by macrophages and ultimately aid tumor cells in entering the bloodstream [5,36]. Once in the vascular system, tumor cells become micrometastases or cells that spread from the primary tumor to surrounding tissues, or distant organs, which represent the main cause of tumor lethality [37,38]. When micrometastases invade another tissue or organ, they can start forming secondary tumors or metastases. The process of metastasis formation is supported by the fact that the disseminated tumor cells mainly represent cell death- and therapy-resistant clones of the primary tumor, giving them a substantial survival advantage compared to surrounding cells [9]. As previously mentioned, such cells have a significantly upregulated PI3K-AKT pathway that facilitates their resistance to apoptosis. The upregulation of the PI3K-AKT pathway represents a consequence of strong signaling from GFs such as IGF-I and EGF-like ligands [5,9]. Moreover, the secretion of factors such as Colony-Stimulating Factor 1 (CSF1) by tumor cells and surrounding macrophages, as well as NRGs, can significantly influence tumor growth and resistance [5,39].
Finally, to gain access to nutrients and oxygen, as well as to proliferate and survive, both primary tumors and metastases can facilitate their vascularization or angiogenesis [40]. De novo blood vessel formation is supported by the recruitment of mature endothelial cells and bone marrow-derived endothelial progenitor cells (EPCs) [25]. GFs such as VEGFs, FGFs, and TGF-β play crucial roles in this process. Moreover, the secretion of proangiogenic GFs by EPCs is important in mediating the angiogenic switch, which is when the antagonism between proangiogenic and antiangiogenic factors switches towards a proangiogenic result [5,41,42]. One of the hallmarks of cancer progression is hypoxia, which contributes to the plasticity and heterogeneity of tumors. Most solid tumors contain a hypoxic core, due to rapid cellular growth that outgrows the blood supply. In these circumstances, hypoxia-inducible factors (HIFs) play a central role in the adaptation of tumor cells to their new environment, dramatically reshaping their transcriptional profile. HIF signaling is modulated by a variety of factors including hormones and GFs which activate signaling pathways that enhance tumor growth and metastatic potential and impair the response to therapeutic agents.
In this review, we will describe the importance of hormones and GFs during cancer onset and progression with a particular focus on hypoxia and the interplay with HIF proteins. We will also discuss how hypoxia influences the efficacy of cancer immunotherapy. Indeed, hypoxia is considered a determinant of the immune-excluded phenotype and a major hindrance to the success of adoptive cell therapies (ACTs).

2. Hypoxia-Inducible Factors

In addition to biological macromolecules such as hormones and GFs, physiological factors are also intimately linked to tumor progression by affecting the TME. One physiological factor that critically influences the TME is the level of oxygen tension (a.k.a. partial oxygen pressure). Under normal physiological conditions, the oxygen tension ranges from 8 to 100 mmHg [43], a state that is referred to as physioxia. When the demand for oxygen exceeds the environmental supply, tissues are described as being in a state of hypoxia. While hypoxia is a normal occurrence during development (e.g., during mammalian embryogenesis), it is also associated with medical conditions [44,45,46,47,48]. Hypoxic areas are present in most solid tumors because of cellular proliferation outgrowing the blood supply [49,50,51], and solid tumors need to become angiogenic to grow beyond 1–2 mm in diameter [52,53,54,55]. Transient acute hypoxia and chronic hypoxia lead to a metabolically heterogeneous TME [56,57,58,59,60,61,62], which creates a strong selective pressure on cells, favoring the growth of more aggressive tumor clones. Therefore, hypoxia is generally clinically associated with poor prognosis across multiple tumor types and is also one of the main causes of therapeutic resistance [59,60,63,64,65,66].
Cellular homeostasis is maintained under hypoxic conditions by a family of heterodimeric transcription factors known as hypoxia-inducible factors (HIFs). HIF proteins contain alpha (HIF-α) and beta (HIF-β or ARNT) subunits and exist as three isoforms (HIF-1, HIF-2, and HIF-3). HIF-1α and HIF-2α are structurally similar except for their transactivation domains. HIF-1α generally binds hypoxia response elements (HREs) close to gene promoters, while HIF-2α targets transcriptional enhancers [67,68,69,70,71,72]. This may explain why they have both common and unique target genes despite binding identical HREs. While HIF-1 plays a major role in the regulation of genes involved in glycolysis, HIF-2 is mainly involved in pluripotent stem cell maintenance and angiogenesis, enhancing the pro-tumorigenic phenotype [73,74,75,76,77]. Finally, HIF-3α lacks a transactivation domain, suggesting that it may suppress the other HIF isoforms [78,79,80,81].

2.1. HIF-Dependent Regulatory Mechanisms

HIF proteins bind to canonical DNA sequences known as HREs, activating the expression of genes involved in a plethora of cellular processes [67,82,83,84,85,86,87,88,89,90,91,92,93]. While HIF proteins clearly exert their functions in a transcription-dependent manner, HIF proteins also impact important cellular processes in manners that are independent of transcription. For example, HIF-1α has been shown to repress DNA replication and induce cell cycle arrest in response to hypoxia independent of transcriptional regulation [94]. Thus, HIFs are the main drivers of cancer progression, exerting their functions through transcription-dependent and -independent mechanisms.

2.2. Oxygen-Dependent Regulation of HIF Signaling

As HIF signaling pathways can have significant impacts on cell function, the production and activity of HIF proteins are tightly regulated. HIF signaling is regulated through numerous cellular mechanisms. Some mechanisms of regulating HIF signaling are oxygen-dependent (e.g., responsive to changes in oxygen tension), while others are oxygen-independent.
HIF activity is largely regulated in an oxygen-dependent manner, such that HIF is active under hypoxia but inactivated in response to normal oxygen tensions. Under physioxia, HIF-1α is produced but is rapidly degraded, resulting in minimal levels of detectable HIF-1α protein. HIF-1α degradation is mediated by prolyl hydrolases (PHDs) which hydroxylate two proline residues within the oxygen-dependent degradation domain of HIF-1α [95,96,97,98] (Figure 2). Hydroxylated HIF-1α is polyubiquitinated by the von Hippel-Lindau tumor suppressor (pVHL), leading to HIF-1α degradation by the 26S proteasome [95,96,97,98]. The factor inhibiting HIF (FIH) protein provides an added layer of HIF repression, suppressing HIF-1 transcription in an oxygen-dependent manner by preventing the recruitment of co-activators [99]. In contrast, under hypoxic conditions, PHDs are inhibited which stabilizes HIF-1α, permitting its nuclear translocation and dimerization with constitutively active HIF-1β. This HIF-1α/β heterodimer then binds to HREs to activate transcription of hypoxia-responsive genes including those involved in promoting glycolytic metabolism, angiogenesis, and survival [67,68,83,100,101,102,103]. Abnormal HIF-1α and consequent upregulation of its target genes have been shown to occur in a wide range of solid tumors as they progress to malignancy [104].
HIF signaling is also regulated in oxygen-independent manners through mechanisms including transition metals (TMs), reactive oxygen species (ROS), reactive nitrogen species (RNS), and mechanical forces. In the late 1980s, a study discovered that TMs could induce hypoxia-like conditions, demonstrating increased erythropoietin gene expression following exposure to TMs such as Co(II), Ni(II), and Mn(II) [105]. Follow-up studies investigating the mechanisms underlying the regulation of erythropoietin gene expression led to the discovery of HIF-1 [68]. It was later found that exposure of cells to various TMs such as Ni(II), Co(II) [106,107], As(III), Cr(VI), and V(V) [108,109] stabilize HIF-1α. This is thought to occur through the inhibition of HIF-1α hydroxylation via two independent mechanisms: (1) substitution of iron by metal ions or (2) iron oxidation in the hydrolases [110,111]. Recent studies provide convincing support for the second hypothesized mechanism. Kaczmarek et al. detected HIF-1α stabilization in human lung epithelial cells in response to exposure to metal and metalloid ions, including some that cannot be used to substitute for iron in the hydroxylases [111]. Additional papers supported that HIF-1α can be stabilized by metal anions that are not iron substitutes (e.g., As(III), Cr(VI), and V(V)) [108,109].
In addition to TMs, reactive molecules such as ROS and RNS have also been shown to stabilize HIFs in a manner that is oxygen-independent. ROS have been discovered to promote HIF-1α transcription and translation in both hypoxia and normoxia [112,113]. For example, superoxide (SO) and hydrogen peroxide have been implicated in ROS-mediated HIF-1α stabilization [113,114,115,116]. Mechanistically, SO has been shown to inhibit VHL from binding to HIF-1α and thus inhibiting PHD activity [115,117] but the mechanisms underlying hydrogen peroxide-mediated HIF stabilization remain unknown. With regard to RNS, increased nitric oxide (NO) leads to HIF-1α accumulation and enhanced DNA binding activity in normoxic conditions [118,119]. This is thought to be mediated by HIF-1α post-translational modifications such as S-nitrosylation [120,121] or by inhibiting PHD activity [122]. The ability of ROS and RNS to mimic hypoxia in this manner suggests that ROS/RNS and hypoxia use similar mechanisms to stabilize HIF-1α [123]. The effects of ROS and RNS on HIF stability have been extensively and elegantly reviewed by Movafagh et al. [124].
Another mechanism triggering HIF stabilization independent of changes in oxygen levels is mechanical force. For example, VEGF-A expression can be induced under non-hypoxic conditions in response to the mechanical stress of the myocardium [125]. A study examining mechanical stress on rat hearts revealed stress-induced VEGF-A induction is preceded by HIF-1α accumulation in the nuclei of cardiac myocytes in a manner that depends on stretch-activated channels and the PI3K/Akt/FRAP pathway that is known to induce HIF-1α [126]. Similar studies revealed that HIF-1α is induced upon the stretching of rat vascular smooth muscle cells [127], and HIF-1α and HIF-2α are induced in skeletal muscle microvascular endothelial cells upon stretching of rat muscles, suggesting that both HIF-1α and HIF-2α contribute to stretch-induced capillary growth under non-hypoxic conditions [128].
Taken together, these findings suggest that HIF signaling is regulated by a variety of oxygen-dependent and -independent mechanisms. Several additional mechanisms are emerging as key regulators of HIF signaling including regulation of HIF by GFs and hormones. In the next section, we will describe the mechanisms through which hormones and GFs impact HIF signaling.

3. Modulation of HIF Signaling by Hormones and GFs

Cancer cell proliferation is stimulated by a variety of hormones such as steroid hormones (SHs) as well as GFs. Likewise, SH and GF axes can also synergistically activate intracellular pathways that are involved in tumor initiation, progression, survival, and resistance to chemotherapy [129,130,131,132,133,134].
Estrogens are SHs that regulate the reproductive process, bone density, and brain function, as well as the endocrine, cardiovascular, and metabolic systems [129,135]. However, these steroids are also considered a risk factor for many types of malignancies such as breast, endometrial, ovarian, prostate, and thyroid tumors [136]. Estrogens mediate numerous biological processes mainly by binding to cognate intracellular receptors, namely Erα and Erβ, which function as ligand-activated transcription factors that induce the transcription of various target genes. In addition, acting through membrane-located Ers or additional estrogen-binding proteins, estrogens elicit rapid responses such as an increase in calcium and nitric oxide levels and the activation of certain kinases [136,137,138,139]. In this regard, several studies have suggested that the G-protein estrogen receptor (GPER) may mediate rapid estrogen actions in both normal and tumor contexts [140,141]. In relation to the pro-tumorigenic properties of estrogens, Erα activates genes involved in the promotion of cell proliferation and cell cycle progression [142,143], inhibition of apoptosis [144], stimulation of angiogenesis [145,146,147], migration and invasion [148], and EMT [149,150]. Beyond Erα, estrogen-induced GPER activation initiates intracellular signaling pathways and triggers changes in gene expression via heterotrimeric G proteins, thus impacting cancer growth, invasion, and metastasis [141,151,152]. Interestingly, GPER can also regulate the expression of pro-tumorigenic mediators such as inflammatory cytokines and angiogenic factors within the TME [153,154,155,156,157]. Of note, Erα and GPER mediate tumor-promoting responses to estrogens by engaging in functional interactions with GF tyrosine kinase receptors including EGFR, IGF-IR, and Fibroblast GF Receptor (FGFR) [132,158,159,160,161,162].
GFR activation represents one of the most common events in the development of human cancers, as GFR activity is implicated in tumor cell proliferation, survival, transformation, angiogenesis, and metastasis [163,164,165,166,167]. EGF, IGFs, FGF, and platelet-derived growth factor (PDGF) are some of the most well-studied GFs that activate numerous pro-carcinogenic signaling pathways including ERK/MAPK, JNK, PI3K/AKT/mTOR, STAT, and PKC transduction cascades [164,166,167,168]. Remarkably, GF/GFRs are attractive therapeutic targets in human cancers due to their hyperactivation and frequent overexpression in cancer cells [164,165,169,170]. Intriguingly, hypoxic mediators as HIFs have been recognized as important drivers of GFR expression and signaling through both oxygen-dependent and -independent mechanisms [171,172,173,174,175,176]. Indeed, high HIF-1α levels have been associated with the overexpression of both GFs and GFRs in human cancers [177,178].

3.1. Hormone-Dependent Regulation of HIF Expression and Signaling

HIF-mediated SH function is involved in non-hypoxic pathways that enable cancer cells to adapt to altered TMEs [179,180,181,182]. Previous studies have focused on the capacity of estrogen to regulate HIF-1α expression in diverse types of cancer cells including thyroid, ovarian, and breast cancer cells, mainly via the PI3K/AKT signaling cascade [180,183,184]. Accordingly, select ER modulators downregulate VEGF-induced angiogenesis by suppressing HIF-1α/VEGFR2 signaling as well as the activation of AKT and ERK axes in breast cancer cells [185]. Furthermore, estrogens downregulate HIF-2α mRNA and protein levels in ER-positive breast cancer cells [186]. These findings are in line with immunohistochemical data showing that high HIF-2α levels are associated with a better overall survival rate in patients with invasive breast tumors [187]. In addition, estrogens cooperate with hypoxia to regulate genes involved in tumor cell growth and differentiation, angiogenesis, protein transport, metabolism, and apoptosis [188]. For instance, estrogen/Erα and hypoxia/HIF-1 transduction pathways can converge on the modulation of the epigenetic modifier histone demethylase KDM4B [188,189,190], affecting cell cycle progression and growth in ER-positive breast cancer [189,191]. The functional interaction between Erα and HIF-1α transduction pathways in breast cancer has been corroborated by evidence showing that HIF-1α may confer resistance to ER antagonists and may be considered as a transcriptional target of Erα [192]. In this respect, it has been reported that Erα binds to estrogen-responsive elements (EREs) located within the HIF-1α gene, thereby enhancing HIF-1α transcriptional activation [192]. The importance of these findings is highlighted by clinical studies revealing that the expression of HIF-1α or a hypoxia metagene signature in ERα-positive breast tumors is associated with aggressive phenotypes and a poor response to endocrine treatment [193,194,195].
The effects of estrogen may be mediated by additional factors, many of which can synergize with and/or regulate HIFs to promote aggressive neoplastic features. For instance, the ERβ variant ERβ2, which is overexpressed in many tumors and associated with decreased overall survival, was shown to stabilize HIF-1α protein levels and promote a hypoxic gene signature in prostate cancer cells and proliferative and invasive phenotypes in triple negative breast cancer cells [196,197]. Moreover, Estrogen-Related Receptor Alpha (ERRα) can interact with HIF-1α, thereby inhibiting its ubiquitination and degradation to promote the adaptation of prostate cancer cells to hypoxic conditions [198]. Further supporting a relationship between estrogens and HIF-1α, our previous studies have demonstrated that estrogenic GPER signaling triggers VEGF expression by upregulating HIF-1α in normoxic breast cancer cells, cancer-associated fibroblasts (CAFs), and mouse xenografts of breast cancer, leading to neoangiogenesis and enhanced tumor growth [155].

3.2. GF-Mediated Regulation of HIF Expression and Signaling

A variety of GFs can activate HIF-1α/VEGF signaling in cancer cells. For instance, it has been shown that under normoxic conditions, activation of the EGFR pathway may lead to increased expression levels and transcriptional activity of HIF-1α via the PI3K/AKT cascade [199]. These findings are in line with data suggesting that increased PI3K activity can promote HIF-1α overexpression in several types of human cancers [65]. EGF was shown to induce the expression of CXC Chemokine Receptor 4 (CXCR4) and promote migration in normoxic non-small cell lung cancer (NSCLC) cells via HIF-1α and the PI3K/AKT and mTOR pathways [200]. Oxygen-independent regulation of HIF-1α through PI3K/AKT and rapamycin has also been shown to occur in breast cancer cells upon treatment with the ErbB3/ErbB4 ligand heregulin, which stimulates HIF-1α synthesis and VEGF-A expression [174].
In breast cancer cells, HIF-1α is required for upregulation and secretion of the pro-tumorigenic cytokine Stem Cell Factor (SCF) by EGF [201]. Likewise, HIF-1α-dependent phosphorylation of STAT3 by EGF stimulates proliferative and invasive responses in colorectal cancer cells [202]. Similarly, EGF-induced expression of the anti-apoptotic protein surviving occurs through HIF-1α, promoting resistance to apoptotic stimuli in breast cancer cells [203].
In addition to HIF-1α regulation by the EGF/EGFR transduction pathway in a hypoxia-independent manner, HIF proteins can regulate their own production through autocrine and additive loops by which hypoxia induces EGF/EGFR expression while the ensuing EGFR signaling synergizes with hypoxia to induce HIF-1α and/or HIF-2α to promote cancer cell survival, migration, and invasion [199,200,204,205,206]. For instance, hypoxia-induced EGFR activation promotes a migratory phenotype in head and neck squamous cell carcinoma (HNSCC) through both HIF-2α and its target gene TGF-α, which may act as an EGFR ligand [207].
Considering that one of the major effects of the EGFR/HIF loop is the induction of VEGF-A in both normoxic and hypoxic conditions [174,208,209], it is not surprising that EGFR-targeting agents exhibit antiangiogenic activity. Among these agents, the monoclonal antibody cetuximab (C225) decreased HIF-1α levels and VEGF-A production in in vitro and in vivo tumor models [210,211,212]. Of note, these effects contributed to vascular normalization and apoptosis-dependent regression of established tumors [212]. In line with these findings, overexpression of HIF-1α conferred resistance to cetuximab-dependent apoptosis and inhibition of VEGF-A secretion in several cancer cells, whereas HIF-1α silencing restored the sensitivity to cetuximab, leading to antitumor responses [213]. Moreover, downregulation of HIF-1α in cancer cells was required for cetuximab-mediated autophagy as well as the sensitivity of HNSCC cells to radiation [214,215]. Similarly, gefitinib and erlotinib reduced vessel formation, decreased vascular permeability, and improved tumor oxygenation in xenograft models through the inhibition of HIF-1α and the production of angiogenic factors such as VEGF-A and IL-8 by tumor cells [216,217,218]. These findings have important implications in cancer therapies employing EGFR inhibitors since resistance to EGFR-directed therapies can be rescued by simultaneously targeting HIF-1α or VEGF-A (reviewed in [209]).
Several lines of evidence suggest that HIF protein accumulation and its activity can occur in a variety of cancer cells in response to IGF-I stimulation [175,219,220,221]. For instance, the exposure of colon carcinoma cells to IGF-I induced HIF-1α protein expression and upregulated VEGF mRNA levels in a PI3K- and MAPK-dependent manner [175]. Accordingly, the involvement of MAPK signaling in the regulation of HIF-1α by IGF-I was observed in breast cancer cells using MEK1/2 inhibitors [222]. Moreover, the engagement of PI3K/AKT and MAPK by IGF-I signaling stimulated HIF-1α and its target genes GPER and VEGF-A in breast cancer cells and CAFs and promoted tumor angiogenesis [223]. In line with these findings, inhibition of the HIF-1α/VEGF axis by both synthetic and natural chemicals such as bisphosphonates, farnesyltransferase inhibitors, epigallocatechin-3-gallate, and dauricine suppressed IGF-I-induced angiogenesis in breast cancer, lung cancer, and HNSCC cells [219,224,225,226]. Of note, in Kaposi sarcoma cells, IFG-I-mediated induction of HIF-1α, HIF-2α, and VEGF-A was lessened by inhibiting IGF-IR, leading to decreased tumor vascularization [220]. Considering that hypoxia promotes HIF-1α-mediated transcription and secretion of IGF-II [227], it is perhaps unsurprising that HIF-1α inhibitors decrease IGF-II signaling and trigger anti-migratory effects in breast cancer cells [228]. These findings might provide new perspectives to clinical studies since IGF-IR inhibitors are non-therapeutic when used alone in cancer patients [229]. Considering that compensatory signals mediated by insulin receptor-A or IGF-IR/IR hybrids may contribute to the lack of efficacy of anti-IGF agents [229], simultaneous inhibition of the IGF system and HIF-1α might prevent compensatory signaling, therefore providing an alternative strategy for the therapeutic management of diverse tumors including breast cancer.
In addition to the EGF and IGF systems, PDGF and basic FGF (bFGF) act as survival factors in cancer cells by engaging stimulatory signaling mechanisms mediated by HIF-1α. For instance, bFGF triggers HIF-1α activation and VEGF-A release via the PI3K/AKT and MEK transduction pathways in normoxic breast cancer cells [176,230,231]. PDGF-BB has been shown to upregulate the Bcl-2 family member named Myeloid Cell Leukemia-1 (Mcl-1) by forming a transcription complex between β-catenin and HIF-1α in prostate cancer cells [232]. Blocking the PDGF system by imatinib inhibits HIF-1α and IGF-I expression in prostate cancer cells and xenograft models [233]. Moreover, HIF-1α promoted lymphatic metastasis of hypoxic breast cancer cells by activating PDGF-B transcription [234], while an AKT/HIF-1α/PDGF-BB autocrine loop mediates hypoxia-induced chemoresistance in liver cancer cells [235].
Taken together, these findings suggest that HIF stabilization and/or activation may be a prerequisite for changes in gene expression and activation of signaling pathways triggered by SHs and GFs (Table 1). These events may promote tumor growth and spread as well as altered responses to therapeutic agents. The next section will discuss the relationship between hypoxia and the immune-excluded phenotype.

4. The Contribution of Hypoxia to the Immune-Excluded Phenotype

During the last decade, significant advances were made in relation to cancer immunotherapy, and it is currently considered a promising therapeutic strategy for many types of cancer. However, despite the proven efficacy of immunotherapy treatments in treating hematological malignancies, solid tumors remain a challenge to treat [236,237,238,239,240]. A main challenge to overcome is the infiltration of T-cells in the tumor core, which is often hypoxic. Several mechanisms have been suggested to explain the immune-excluded phenotype; however, an integrated understanding of the role played by various determinants of immune exclusion is still lacking. Hypoxia is a hallmark of most solid tumors and one of the most relevant factors involved in the immune excluded phenotype [49,50,51,241]. Hypoxia is responsible for shaping the TME in a unique way, affecting both gene transcription and chromatin remodeling in cancer and immune cells and therefore plays a role in the formation of mechanical and functional barriers. Indeed, hypoxia is a major suppressor of the immune system that alters the expression of cytokines, inducing the expression of co-inhibitory ligands and recruiting immune-suppressive cell populations [241,242,243,244,245].

4.1. Physical Barriers

Cancer cells produce pro-angiogenic factors through upregulation of HIF-induced genes (Table 2), which are responsible for the spatial-temporal heterogeneity in blood flow observed in solid tumors [246,247]. Indeed, the new blood vessels are often abnormal in that they are leaky and have an aberrant structure, which further enhances the hypoxic conditions [55,248,249,250]. Consequently, reduced blood flow, acidosis, and increased interstitial fluid pressure affect the fitness and infiltration of T-cells in the tumor core, which leads to a poor clinical response to immunotherapy [243,251,252]. For example, Jayaprakash et al. (2018) showed that in preclinical prostate cancer, hypoxic zones resisted T-cell infiltration even in the context of Cytotoxic T-Lymphocyte-Associated Protein 4 (CTLA-4) and Programmed Cell Death Protein 1 (PD-1) blockade. However, treating these tumors with the hypoxia-activated prodrug TH-302 in combination with checkpoint blockade cured more than 80% of tumors in a mouse model [243].
HIF-1 upregulates VEGF-A, which is a main pro-angiogenic factor that is produced by fibroblasts and endothelial cells [292,293]. VEGF is also responsible for ECM fibrosis in cancer as it induces the secretion of fibronectin and collagen type-I by stromal cells, activated resident fibroblasts, and attracted fibroblasts [357,358]. Hypoxia-dependent increases in tumor stroma density are a common occurrence in cancer and are due to the increased secretion of fibrous material and collagen-modifying enzymes [253,254,255,256,359]. Depletion of HIF-1, but not HIF-2, inhibited collagen deposition in vitro and decreased stromal density in orthotopic tumor [257]. However, collagen degradation in certain regions of the TME has also been reported, due to the activity of matrix metalloproteinases (MMPs) [360,361]. Collagen degradation is an important step during TME remodeling and HIF-1 has been associated with transcriptional upregulation of MMP2 and MMP9 in vitro [258,362]. The interplay of these processes, especially the increased stromal stiffness in tumors, constitutes an important barrier against T-cell accessibility. This is caused by mechanical constraints and decreased bioavailability of signaling. Kuczek et al. (2019) cultured T-cells and human breast cancer cells in 3D conditions with high or low collagen density. While the proliferation of cancer cells was unaffected in high collagen density matrices, T-cell proliferation was significantly decreased, along with a higher ratio of CD4+ to CD8+ cells and downregulation of cytotoxic markers [363]. Recruitment and activation of fibroblasts are also mediated by a variety of cytokines (Table 2) including TGF-β that is secreted by cancer cells and other stroma cells. TGF-β transforms fibroblasts into CAFs, which are responsible for the secretion of fibrous material and cytokines [259,364,365]. CAFs play various roles in the TME, they are morphologically different from fibroblasts, and they possess an enhanced migratory potential and proliferative capacity. In hypoxia, a positive feedback loop has been reported between TGF-β and HIF proteins [366]. Moreover, HIF-1 activates the TGF-β/Smad signaling cascade leading to increased collagen deposition in dermal fibroblasts [367].
CAFs promote functional barriers that contribute to angiogenesis, tumor growth, metastasis, immune suppression, and metabolic reprogramming through the secretion of chemokines such as CXCL12. CXCL12 is also involved in the EMT in various tumor types [368,369]. Hypoxia-induced cytokines secreted by cancer cells, myeloid cells, and mesenchymal cells are involved in the EMT (Table 2). Such cytokines include Tumor Necrosis Factor Alpha (TNF-α), TGF-β, Interleukin-1 (IL-1), IL-6, and IL-8 [272,273,274,275,276,370,371]. Moreover, it has been shown that EGF, hepatocyte growth factor (HGF), bFGF, and PDGF are also involved in the induction of transcription factors responsible for EMT progression [277,278,372,373]. EMT signatures are inversely correlated with T-cell infiltration in NSCLC, urothelial cancer and with resistance to PD-1 blockade in melanoma [279,374,375,376]. However, a few recent studies showed a positive correlation between T-cell infiltration and stromal EMT-related gene expression, in various malignancies [377,378]. Overall, the TME contains a variety of physical barriers that act synergistically with functional barriers to create an immunosuppressive environment and impede T-cell activity.

4.2. Functional Barriers

Functional barriers embody a class of impediments that impair T-cell activity despite their interaction with cancer cells and include metabolic barriers such as nutrient deprivation in the TME, danger sensing pathways, soluble factors, and cell-intrinsic signaling (Table 2). These determinants affect T-cell penetration and expansion in the tumor nests.
Hypoxic conditions lead to a shift from oxidative to glycolytic metabolism due to the upregulation of HIF-related genes. This phenomenon is called the Warburg effect and it is almost universal in cancers even when oxygen levels are comprised in a physiological range, despite also occurring physiologically [379,380,381]. HIF-1 regulates enzymes responsible for glucose transport (GLUTs) to maintain cellular ATP pools [299,300]. In solid tumors, GLUT upregulation is usually correlated with poor prognosis [382]. HIF-1 is also responsible for the upregulation of enzymes involved in glycolysis and pH regulation [309,383,384]. Enhanced glycolysis leads to increased amounts of lactate in the TME, which is responsible for the acidification of the ECM and subsequent decreases in T- and NK cell function and survival [385,386]. Furthermore, hypoxia induces carbonic anhydrases (Cas), Na+/H+ exchanger (NHE1), bicarbonate transporters (SLC4A4), and Indoleamine 2,3 Dioxygenase (IDO), which contribute to the acidification and tryptophan depletion in the TME [310,311,316,387]. In addition to tryptophan, the TME can be deprived of other amino acids essential for T-cell activity such as arginine [388,389]. HIF-1 is also responsible for GLS1 upregulation in cancer cells, which enhances anabolic metabolism through the hydrolyzation of glutamine into glutamate [317,390]. Decreases in the extracellular concentrations of glutamine and leucine inhibit the differentiation of Th1 and Th17 effector lymphocytes while maintaining regulatory T-cell differentiation [391,392]. Finally, elevated concentrations of extracellular potassium are common in solid tumors due to the downregulation of the potassium channel Kv1.3 [318,319]. Ionic imbalance, low pH, low glucose, and insufficient amino acid presence in the TME impair effector T-cell proliferation and cytokine production, leading to T-cell exhaustion.
HIF proteins are responsible for the upregulation and secretion of a variety of chemokines responsible for myeloid cell recruitment (Table 2). In hypoxia, immune-suppressive populations such as regulatory T-cells, myeloid-derived suppressor cells (MDSCs), and tumor-associated macrophages (TAMs) infiltrate the tumor site and create a functional barrier to the infiltration of T- and NK cells [393,394,395]. TAM receptor kinases such as Tyro3, Axl, and Mertk promote the phagocytosis of apoptotic cells, binding to “eat-me” signals which are displayed on apoptotic cell membranes [396]. Several studies have reported that TAM receptors are involved in the hypoxic response and play a variety of roles in cancer cells and tumor-infiltrating immune cells. Different signaling cascades may lead to the regulation of distinct TAM-associated functions [349,350,351]. Despite some evidence suggesting that TAM receptors may be involved in the immune-suppressive phenotype, their function requires further investigation. Moreover, HIF-1 induces the expression of CD47 which is a “don’t eat-me” signal (CD47/signal regulatory protein (SIRP)-α axis) that blocks prophagocytic signals and promotes tumor escape from immune surveillance [397,398,399,400]. Hypoxia also upregulates CD39 and CD73 which drive the shift from a proinflammatory to an anti-inflammatory environment mediated by an accumulation of adenosine in the TME [401,402]. Adenosine receptors on the surface of T-cells are upregulated by HIF-1 and HIF-2, and adenosine intake by T-cells leads to the accumulation of intracellular cyclic adenosine monophosphate (cAMP) [403,404]. cAMP plays an important role in immunosuppression, inhibiting T-cell receptor-induced T-cell activation [405,406,407]. Another functional barrier is the hypoxia-dependent downregulation of proteins necessary for immune cell recognition. For example, hypoxic tumors exhibit downregulation of major histocompatibility class-I (MHC-I) molecules, which impairs their recognition by cytotoxic T-cells (CTLs) [348]. In vivo investigations have revealed that hypoxia triggers the inhibition of IFN-γ–dependent MHC-I upregulation, a phenomenon that is reversible upon re-oxygenation [408].
Several comprehensive reviews have recently commented on the role of hypoxia as a determinant of the immune-excluded phenotype in solid tumors, and this chapter provided an overview of the major mechanisms involved. Immune-excluded tumors are different from homogeneously-infiltrated tumors, as they contain gradients of exclusion. Immune-excluded tumors may occur because of numerous underlying mechanisms and represent a significant challenge for immunotherapy. Gradients of immune infiltration are peculiar within each tumor environment and are likely not present in silent tumors. In cold tumors, a lack of chemo-attraction may trigger a predominant phenotype rather than the presence of barriers. The next section will provide some perspectives on the interplay between hypoxia and immunotherapies.

5. Perspectives on the Interplay between Hypoxia and Immunotherapies

Hypoxia is associated with a poor response to radiotherapy and chemotherapy and, therefore, is considered a negative prognostic indicator [409,410]. Hypoxia promotes tumor progression through angiogenesis, stemness, increased cancer cell survival, and metastasis. In the last few years, hypoxia-activated prodrugs (HAPs) have been tested in clinical trials as strategies to reduce hypoxic areas and improve cancer therapies. HAPs are inactive compounds that are converted to active drugs in hypoxic tissues [411,412]. However, low partial oxygen pressure is a common feature of many healthy tissues, for example in secondary lymphoid organs where the oxygen level is approximately 2.5%. Immune cells require HIF-mediated responses for their effector functions and their inhibition can significantly impair their functionality. Class I HAPs require mild hypoxia for their activation, necessitating the discovery of HAPs that require more severe hypoxia (Class II) to avoid undesired side effects [413]. Clinical trials using HAPs alone reported disappointing results, and studies assessing the efficacy of HAPs combined with immunotherapy in the treatment of solid tumors are currently underway [414,415,416]. Jayaprakash et al. (2018) reported that hypoxic zones were prevalent in preclinical prostate cancer and resisted T-cell infiltration, even in the context of CTLA-4 and PD-1 blockade [243]. Indeed, HIF-1 directly binds the promoter of PD-L1, inducing its transcription in various tumor cells [353,417]. In clear cell renal cell carcinoma, HIF-2 triggers PD-L1 upregulation [418,419]. The PTEN/PI3K pathways are responsible for HIF-1-mediated upregulation of CTLA-4, PD-L1, and HLA-G checkpoint molecules in several different mice and human tumor cell lines. Jayaprakash et al. (2018) showed that combination therapy with TH-302 HAP and T-cell immune checkpoint blockades (CTLA-4 and PD-1) cured more than 80% of tumors in a mouse prostate-derived TRAMP-C2 model. It is thought that the combination of the two treatments increased T-cell infiltration into hypoxic areas due to improved vascularization. Moreover, both myeloid-derived suppressor cells and granulocytic subsets were reduced in the TME. However, the results were less conclusive in spontaneous TRAMP models than in the ectopic models [243].
Approaches to improve the efficacy of immunotherapies in hypoxic tumors might include improving the function of the existing microvascular network. As previously mentioned, the ECM is often very dense in solid tumors and this may increase the pressure on the vasculature, leading to its collapse [420,421,422]. Another potential strategy may be the normalization of abnormal blood vessels produced in hypoxic conditions through metronomic dosing of VEGFs inhibitors [421,423,424]. Both approaches may be useful to increase oxygenation in the tissues and assist in chimeric antigen receptor (CAR) T-cell delivery. Two clinical trials (NCT03634332 and NCT02563548) are currently assessing the efficacy of a combination of PEGPH20 and pembrolizumab for patients with “Hyaluronan High” solid tumors. PEGPH20 is a pegylated recombinant human hyaluronidase that enzymatically degrades hyaluronan in the tumor stroma. Hyaluronan is a major component of the stroma, and the purpose of these clinical trials was to reduce the stromal density to improve the function of the vascular network and facilitate pembrolizumab penetration. However, a clinical trial (NCT01839487) reported that PEGPH20 in combination with nab-paclitaxel and gemcitabine did not provide any benefit to patients with Stage IV pancreatic ductal adenocarcinoma.
A major issue associated with ACT is the off-target toxicity, and recent studies indicated that hypoxia may be exploited to obtain more precise CAR T therapies. A hypoxia-inducible transcription amplification system (HiTAsystem) was developed to control CAR expression in T-cells (HiTA-CAR-T) [425]. An anti-Her2 CAR was expressed only in hypoxic environments, and HiTA-CAR T showed significant hypoxia-dependent tumor suppression in murine xenograft models. Moreover, no toxicity to livers expressing human target antigens in naïve mice was detected. Kosti et al. (2021) also developed a hypoxia-inducible CAR construct in which the transcription was mediated by HIF-1 (HypoxiCAR) [426]. The expression of a pan-ErbB-targeted CAR within hypoxic solid tumors demonstrated anti-tumor efficacy without off-tumor toxicity in murine xenografts models. Interestingly, they also reported that HypoxiCAR T-cells were not excluded from HIF-1-stabilized regions of the tumor (Figure 3).
During the last decade, several studies have reported that hypoxia may be beneficial for CAR T-cell activation, leading to increased T-cell effector function. Ectopic HIF expression has also been used to improve the antitumor efficacy of CD8+ T-cells and may be an interesting strategy to apply to ACT cancer therapy. For example, Gropper et al. (2017) showed that CTLs activated under hypoxic conditions (1% oxygen) have improved intrinsic cytotoxicity rather than migration potential inside tumors in vitro [427]. CTLs had increased amounts of granzyme-B in each granule, leading to improved mouse survival. However, hypoxia inhibited CD8+ T-cell proliferation in culture and increased exhaustion markers such as T-cell immunoglobulin, Mucin-Domain-Containing 3 (TIM3), and Lymphocyte Activation Gene 3 (LAG3). Conversely, CTLA4 and PD1 expression remained unaltered in hypoxia. Caldwell et al. (2001) performed similar experiments in CTLs at 2.5% oxygen and found an increase in effector functions and cell survival and a decrease in cytokine secretion and T-cell expansion [428]. Other reports indicated that 1% oxygen did not modify CTL cytotoxicity but induced IL-10 secretion and upregulation of CD137 and CD25 [429]. Roman et al. (2010) performed a similar study on CD4+ T-cells and found that stimulation under hypoxic conditions increased the secretion of effector cytokines, especially IFN-γ [430]. Deletion of the transcription factor Nuclear Factor Erythroid 2-Related Factor 2 (Nrf2) impaired the enhancing effect of hypoxia. Another study showed that HIF-2α, but not HIF-1α, drove broad transcriptional changes in CD8+ T-cells, resulting in increased cytotoxic differentiation and cytolytic function against tumors [431]. Moreover, a HIF-2 form that is insensitive to FIH was delivered with anti-CD19 CAR T-cells. These CAR T-cells showed enhanced cytolytic function in vitro and in a B-cell lymphoma xenograft ACT mouse model. Finally, genetic deletion of VHL and genes encoding for PHDs caused the accumulation of HIF-1 and HIF-2 in T-cells, enhancing their cytotoxic differentiation, and improving the rejection of primary and metastatic melanoma tumors [431].
Finally, Berahovich et al. (2019) characterized anti-CD19 and anti-BCMA human CAR T-cells generated at 18% and 1% oxygen [432]. Under hypoxia, CAR T-cells were associated with reduced proliferation and a less differentiated phenotype (higher CD4:CD8 ratio). Cytolytic activity and PD-1 upregulation was found to be similar in hypoxic and normoxic CAR T-cells. However, cytokine production and granzyme release were significantly decreased in hypoxia. It seems that oxygen tension may play a role during CAR T-cell differentiation, as has been previously reported [432]. The Berahovich et al. (2019) study found that hypoxia favored the differentiation in central memory T-cells (Tcm) compared to effector memory T-cells (Tem). The increase in central memory CAR T, and in less differentiated CAR T-cells in general, may be favorable and improve the efficacy of therapeutic treatments, as reported in previous clinical trials [432].
This observation regarding the prevalence of Tcm in hypoxia contrasts with findings reported by other studies, in which T-cell activation in hypoxia leads to increased differentiation of T-cells towards more lytic effector cells and impairs the proliferation of naïve and Tcm [428,429,433,434]. Indeed, Tem mainly rely on glycolysis instead of oxidative phosphorylation to support their metabolism, which is more favorable under hypoxic conditions. The inconsistencies reported across different studies may be explained by differences in the experimental protocols adopted, the type of stimulus used for activation, and the type of cells used. The timing of activation and analysis may also play a role, as HIF-1 is mainly stabilized during the early stages of hypoxia while HIF-2 is stabilized during the later stages and plays a main role in chronic hypoxia.

6. Conclusions

Hormones and GFs are emerging as key regulators of HIF signaling, with impacts on tumor growth, metastasis, and response to therapeutic agents. HIF proteins can thus be viewed as “central stations” that respond to various facets of the TME and facilitate the adaptation of tumor cells to their environment. Hormones and growth factors can shift the tumor toward a malignant phenotype, due to their influence on HIF proteins. Additional studies examining the connection between hormones, GFs, and hypoxia are vital, as the latter is a major determinant of the immune excluded phenotype. Furthering our understanding of the relationship between hypoxia and immune exclusion will indeed provide key insight into the development of novel and effective ACTs.

Author Contributions

Writing, review and editing: All authors contributed equally. All authors have read and agreed to the published version of the manuscript.

Funding

M.M. was supported by Fondazione AIRC (IG 21322). R.L. and M.M. acknowledge (i) the special award, namely, “Department of Excellence 2018–2022” (Italian Law 232/2016) to the Department of Pharmacy, Health and Nutritional Sciences of the University of Calabria (Italy), (ii) the “Sistema Integrato di Laboratori per L’Ambiente—(SILA) PONa3_00341”.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ACTAdoptive cell therapy
bFGFBasic FGF
CAFsCancer-associated fibroblasts
CAsCarbonic anhydrases
TcmCentral memory T cells
CSF1Colony-Stimulating Factor 1
CXCR4CXC chemokine receptor 4
cAMPCyclic adenosine monophosphate
CTLsCytotoxic T cells
CTLA-4Cytotoxic T-lymphocyte associated protein 4
DCISDuctal carcinoma in situ
TemEffector memory T cells
EGFREGF receptor
EPCsEndothelial progenitor cells
EGFsEpidermal Growth Factors
EMTEpithelial-mesenchymal transition
EREstrogen Receptor
EREsEstrogen responsive elements
ERRαEstrogen-related receptor-α
ECMExtracellular matrix
FIHFactor inhibiting HIF
FIHFactors Inhibiting HIF
FGFFibroblast growth factor
FGFRFibroblast growth factor receptor
GFRsGF receptors
GLUTsGlucose transport
GPERG-protein estrogen receptor
GFsGrowth factors
GHGrowth Hormone
GHRGrowth Hormone Receptor
HGFHepatocyte growth factor
HAPsHypoxia-activated prodrugs
HIFsHypoxia-inducible factors
HiTAsystemHypoxia-inducible transcription amplification system
HREsHypoxia-regulated elements
IDOIndoleamine 2,3 dioxygenase
IGF-IRInsulin-like growth factor 1 receptor
IGFsInsulin-like Growth Factors
IL-Interleukin
KSKaposi sarcoma
LAG3Lymphocyte-activation gene 3
MHC-IMajor histocompatibility class-I
MMPsMatrix metalloproteinases
mAbsMonoclonal antibodies
TIM3Mucin-domain containing-3
Mcl-1Myeloid cell leukemia-1
MDSCsMyeloid-derived suppressor cells
HNSCCNeck squamous cell carcinoma
NRGsNeuregulins
NONitric oxide
NSCLCNon-small cell lung cancer
Nrf2Nuclear erythroid 2 p45–related factor 2
PDGFPlatelet-derived growth factor
PRsProgesterone Receptors
PD-1Programmed cell death protein 1
PHDsProlyl hydrolases
PHDsProlyl hydroxylases domains
RNSReactive nitrogen species
ROSReactive oxygen species
STAT3Signal Transducer and Activator of Transcription 3
STATsSignal Transducers and Activator of Transcription
SCFStem cell factor
SHsSteroid ex hormones
SOSuperoxide
TGF-αTransforming Growth Factor-α
TGF-βTransforming Growth Factor-β
TMsTransition metals
TMETumor microenvironment
TNF-αTumor necrosis factor α
TAMsTumor-associated macrophages
TKIsTyrosine kinase inhibitors
VEGFsVascular Endothelial Growth Factors
pVHLVon Hippel-Lindau tumor suppressor

References

  1. Reznikov, A. Hormonal Impact on Tumor Growth and Progression. Exp. Oncol. 2015, 37, 162–172. [Google Scholar] [CrossRef]
  2. Hollinger, J.O.; Alvarez-Urena, P.; Ducheyne, P.; Srinivasan, A.; Baskin, J.; Waters, H.; Gruber, R. Comprehensive Biomaterials II. 6.2 Bone Tissue Engineering: Growth Factors and Cytokines; Elsevier: Amsterdam, The Netherlands, 2017; pp. 20–53. [Google Scholar]
  3. Lee, E.Y.; Parry, G.; Bissell, M.J. Modulation of Secreted Proteins of Mouse Mammary Epithelial Cells by the Collagenous Substrata. J. Cell Biol. 1984, 98, 146–155. [Google Scholar] [CrossRef] [PubMed]
  4. Sherwood, L.M.; Parris, E.E.; Folkman, J. Tumor Angiogenesis: Therapeutic Implications. N. Engl. J. Med. 1971, 285, 1182–1186. [Google Scholar] [CrossRef] [PubMed]
  5. Witsch, E.; Sela, M.; Yarden, Y. Roles for Growth Factors in Cancer Progression. Physiology 2010, 25, 85–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Sporn, M.B.; Todaro, G.J. Autocrine Secretion and Malignant Transformation of Cells. N. Engl. J. Med. 1980, 303, 878–880. [Google Scholar] [CrossRef]
  7. Yeung, S.; Gagel, R.; Kufe, D.; Pollock, R.; Weichselbaum, R. Endocrine Paraneoplastic Syndromes (“ectopic” Hormone Production). In Holland-Frei Cancer Medicine, 6th ed.; BC Decker Inc.: Hamilton, ON, USA, 2003. [Google Scholar]
  8. Hinson, J.; Raven, P.; Chew, S. Receptors and Hormone Action. In The Endocrine System; Elsevier: Amsterdam, The Netherlands, 2010; pp. 15–26. [Google Scholar]
  9. Stone, W.L.; Leavitt, L.; Varacallo, M. Physiology, Growth Factor; StatPearls Publishing: Treasure Island, FL, USA, 2017. [Google Scholar]
  10. Chen, Y.; Takeshita, A.; Ozaki, K.; Kitano, S.; Hanazawa, S. Transcriptional Regulation by Transforming Growth Factor β of the Expression of Retinoic Acid and Retinoid X Receptor Genes in Osteoblastic Cells Is Mediated through AP-1. J. Biol. Chem. 1996, 271, 31602–31606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Beato, M.; Chávez, S.; Truss, M. Transcriptional Regulation by Steroid Hormones. Steroids 1996, 61, 240–251. [Google Scholar] [CrossRef]
  12. Ing, N.H. Steroid Hormones Regulate Gene Expression Posttranscriptionally by Altering the Stabilities of Messenger RNAs. Biol. Reprod. 2005, 72, 1290–1296. [Google Scholar] [CrossRef] [Green Version]
  13. Boguszewski, C.L.; da Silva Boguszewski, M.C. Growth Hormone’s Links to Cancer. Endocr. Rev. 2019, 40, 558–574. [Google Scholar] [CrossRef]
  14. Sherbet, G.V. Hormonal Influences on Cancer Progression and Prognosis. Vitam. Horm. 2005, 71, 147–200. [Google Scholar]
  15. Waters, M.J. The Growth Hormone Receptor. Growth Horm. IGF Res. 2016, 28, 6–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Perry, J.K.; Liu, D.-X.; Wu, Z.-S.; Zhu, T.; Lobie, P.E. Growth Hormone and Cancer: An Update on Progress. Curr. Opin. Endocrinol. Diabetes Obes. 2013, 20, 307–313. [Google Scholar] [CrossRef] [PubMed]
  17. Kang, H.; Wu, Q.; Sun, A.; Liu, X.; Fan, Y.; Deng, X. Cancer Cell Glycocalyx and Its Significance in Cancer Progression. Int. J. Mol. Sci. 2018, 19, 2484. [Google Scholar] [CrossRef] [Green Version]
  18. Miranda, O.; Farooqui, M.; Siegfried, J.M. Status of Agents Targeting the HGF/c-Met Axis in Lung Cancer. Cancers 2018, 10, 280. [Google Scholar] [CrossRef] [Green Version]
  19. Cevenini, A.; Orrù, S.; Mancini, A.; Alfieri, A.; Buono, P.; Imperlini, E. Molecular Signatures of the Insulin-like Growth Factor 1-Mediated Epithelial-Mesenchymal Transition in Breast, Lung and Gastric Cancers. Int. J. Mol. Sci. 2018, 19, 2411. [Google Scholar] [CrossRef] [Green Version]
  20. Greaves, M.; Maley, C.C. Clonal Evolution in Cancer. Nature 2012, 481, 306–313. [Google Scholar] [CrossRef]
  21. Huang, H.-J.S.; Nagane, M.; Klingbeil, C.K.; Lin, H.; Nishikawa, R.; Ji, X.-D.; Huang, C.-M.; Gill, G.N.; Steven Wiley, H.; Cavenee, W.K. The Enhanced Tumorigenic Activity of a Mutant Epidermal Growth Factor Receptor Common in Human Cancers Is Mediated by Threshold Levels of Constitutive Tyrosine Phosphorylation and Unattenuated Signaling. J. Biol. Chem. 1997, 272, 2927–2935. [Google Scholar] [CrossRef] [Green Version]
  22. Almeida, A.; Muleris, M.; Dutrillaux, B.; Malfoy, B. The Insulin-like Growth Factor I Receptor Gene Is the Target for the 15q26 Amplicon in Breast Cancer. Genes Chromosomes Cancer 1994, 11, 63–65. [Google Scholar] [CrossRef] [PubMed]
  23. Arteaga, C.L.; Hurd, S.D.; Winnier, A.R.; Johnson, M.D.; Fendly, B.M.; Forbes, J.T. Anti-Transforming Growth Factor (TGF)-Beta Antibodies Inhibit Breast Cancer Cell Tumorigenicity and Increase Mouse Spleen Natural Killer Cell Activity. Implications for a Possible Role of Tumor Cell/host TGF-Beta Interactions in Human Breast Cancer Progression. J. Clin. Investig. 1993, 92, 2569–2576. [Google Scholar]
  24. Meijnen, P.; Peterse, J.L.; Antonini, N.; Rutgers, E.J.T.; van de Vijver, M.J. Immunohistochemical Categorisation of Ductal Carcinoma in Situ of the Breast. Br. J. Cancer 2008, 98, 137–142. [Google Scholar] [CrossRef] [Green Version]
  25. Asahara, T.; Murohara, T.; Sullivan, A.; Silver, M.; van der Zee, R.; Li, T.; Witzenbichler, B.; Schatteman, G.; Isner, J.M. Isolation of Putative Progenitor Endothelial Cells for Angiogenesis. Science 1997, 275, 964–967. [Google Scholar] [CrossRef]
  26. Yilmaz, M.; Christofori, G. EMT, the Cytoskeleton, and Cancer Cell Invasion. Cancer Metastasis Rev. 2009, 28, 15–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Turner, T.; Chen, P.; Goodly, L.J.; Wells, A. EGF Receptor Signaling Enhances In Vivo Invasiveness of DU-145 Human Prostate Carcinoma Cells. Clin. Exp. Metastasis 1996, 14, 409–418. [Google Scholar] [CrossRef] [PubMed]
  28. Liang, Y.; Zhu, F.; Zhang, H.; Chen, D.; Zhang, X.; Gao, Q.; Li, Y. Conditional Ablation of TGF-β Signaling Inhibits Tumor Progression and Invasion in an Induced Mouse Bladder Cancer Model. Sci. Rep. 2016, 6, 29479. [Google Scholar] [CrossRef] [PubMed]
  29. Stetler-Stevenson, W.G. Matrix Metalloproteinases in Angiogenesis: A Moving Target for Therapeutic Intervention. J. Clin. Investig. 1999, 103, 1237–1241. [Google Scholar] [CrossRef] [Green Version]
  30. Wells, A.; Kassis, J.; Solava, J.; Turner, T.; Lauffenburger, D.A. Growth Factor-Induced Cell Motility in Tumor Invasion. Acta Oncol. 2002, 41, 124–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Roche, J. The Epithelial-to-Mesenchymal Transition in Cancer. Cancers 2018, 10, 52. [Google Scholar] [CrossRef] [Green Version]
  32. Mendonsa, A.M.; Na, T.-Y.; Gumbiner, B.M. E-Cadherin in Contact Inhibition and Cancer. Oncogene 2018, 37, 4769–4780. [Google Scholar] [CrossRef]
  33. Fan, Y.; Shen, B.; Tan, M.; Mu, X.; Qin, Y.; Zhang, F.; Liu, Y. TGF-β–Induced Upregulation of malat1 Promotes Bladder Cancer Metastasis by Associating with suz12. Clin. Cancer Res. 2014, 20, 1531–1541. [Google Scholar] [CrossRef] [Green Version]
  34. Atlas, E.; Cardillo, M.; Mehmi, I.; Zahedkargaran, H. Heregulin Is Sufficient for the Promotion of Tumorigenicity and Metastasis of Breast Cancer Cells In Vivo. Mol. Cancer Res. 2003, 1, 165–175. [Google Scholar]
  35. Katt, M.E.; Wong, A.D.; Searson, P.C. Dissemination from a Solid Tumor: Examining the Multiple Parallel Pathways. Trends Cancer Res. 2018, 4, 20–37. [Google Scholar] [CrossRef]
  36. Padua, D.; Zhang, X.H.-F.; Wang, Q.; Nadal, C.; Gerald, W.L.; Gomis, R.R.; Massagué, J. TGFβ Primes Breast Tumors for Lung Metastasis Seeding through Angiopoietin-like 4. Cell 2008, 133, 66–77. [Google Scholar] [CrossRef] [Green Version]
  37. Tarin, D. Cell and Tissue Interactions in Carcinogenesis and Metastasis and Their Clinical Significance. Semin. Cancer Biol. 2011, 21, 72–82. [Google Scholar] [CrossRef] [PubMed]
  38. Seyfried, T.N.; Huysentruyt, L.C. On the Origin of Cancer Metastasis. Crit. Rev. Oncog. 2013, 18, 43–73. [Google Scholar] [CrossRef] [Green Version]
  39. Aharinejad, S.; Paulus, P.; Sioud, M.; Hofmann, M.; Zins, K.; Schafer, R.; Stanley, E.R.; Abraham, D. Experimental Therapeutics, Molecular Targets, and Chemical Biology-Colony-Stimulating Factor-1 Blockade by Antisense Oligonucleotides and Small Interfering RNAs Suppresses Growth of Human Mammary. Cancer Res. 2004, 64, 5378–5384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Lugano, R.; Ramachandran, M.; Dimberg, A. Tumor Angiogenesis: Causes, Consequences, Challenges and Opportunities. Cell. Mol. Life Sci. 2020, 77, 1745–1770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Lyden, D.; Hattori, K.; Dias, S.; Costa, C.; Blaikie, P.; Butros, L.; Chadburn, A.; Heissig, B.; Marks, W.; Witte, L.; et al. Impaired Recruitment of Bone-Marrow-Derived Endothelial and Hematopoietic Precursor Cells Blocks Tumor Angiogenesis and Growth. Nat. Med. 2001, 7, 1194–1201. [Google Scholar] [CrossRef] [PubMed]
  42. Baeriswyl, V.; Christofori, G. The Angiogenic Switch in Carcinogenesis. Semin. Cancer Biol. 2009, 19, 329–337. [Google Scholar] [CrossRef]
  43. Ortiz-Prado, E.; Dunn, J.F.; Vasconez, J.; Castillo, D.; Viscor, G. Partial Pressure of Oxygen in the Human Body: A General Review. Am. J. Blood Res. 2019, 9, 1–14. [Google Scholar] [PubMed]
  44. Semenza, G.L. Hypoxia-Inducible Factors in Physiology and Medicine. Cell 2012, 148, 399–408. [Google Scholar] [CrossRef] [Green Version]
  45. Hlatky, M.A.; Quertermous, T.; Boothroyd, D.B.; Priest, J.R.; Glassford, A.J.; Myers, R.M.; Fortmann, S.P.; Iribarren, C.; Tabor, H.K.; Assimes, T.L.; et al. Polymorphisms in Hypoxia Inducible Factor 1 and the Initial Clinical Presentation of Coronary Disease. Am. Heart J. 2007, 154, 1035–1042. [Google Scholar] [CrossRef] [PubMed]
  46. Jha, N.K.; Jha, S.K.; Sharma, R.; Kumar, D.; Ambasta, R.K.; Kumar, P. Hypoxia-Induced Signaling Activation in Neurodegenerative Diseases: Targets for New Therapeutic Strategies. J. Alzheimers Dis. 2018, 62, 15–38. [Google Scholar] [CrossRef] [PubMed]
  47. Pang, B.; Zhao, F.; Zhou, Y.; He, B.; Huang, W.; Zhang, F.; Long, Y.-G.; Xia, X.; Liu, M.-L.; Jiang, Y.-H. Systematic Review and Meta-Analysis of the Impact of Hypoxia on Infarcted Myocardium: Better or Worse? Cell. Physiol. Biochem. 2018, 51, 949–960. [Google Scholar] [CrossRef]
  48. Levenson, N.I.; Adolph, R.J.; Romhilt, D.W.; Gabel, M.; Sodd, V.J.; August, L.S. Effects of Myocardial Hypoxia and Ischemia on Myocardial Scintigraphy. Am. J. Cardiol. 1975, 35, 251–257. [Google Scholar] [CrossRef]
  49. Vaupel, P.; Schlenger, K.; Knoop, C.; Höckel, M. Oxygenation of Human Tumors: Evaluation of Tissue Oxygen Distribution in Breast Cancers by Computerized O2 Tension Measurements. Cancer Res. 1991, 51, 3316–3322. [Google Scholar] [PubMed]
  50. Braun, R.D.; Lanzen, J.L.; Snyder, S.A.; Dewhirst, M.W. Comparison of Tumor and Normal Tissue Oxygen Tension Measurements Using OxyLite or Microelectrodes in Rodents. Am. J. Physiol. Heart Circ. Physiol. 2001, 280, H2533–H2544. [Google Scholar] [CrossRef] [PubMed]
  51. McKeown, S.R. Defining Normoxia, Physoxia and Hypoxia in Tumours-Implications for Treatment Response. Br. J. Radiol. 2014, 87, 20130676. [Google Scholar] [CrossRef] [Green Version]
  52. Thomlinson, R.H.; Gray, L.H. The Histological Structure of Some Human Lung Cancers and the Possible Implications for Radiotherapy. Br. J. Cancer 1955, 9, 539–549. [Google Scholar] [CrossRef] [Green Version]
  53. Chapman, J.D. The Detection and Measurement of Hypoxic Cells in Solid Tumors. Cancer 1984, 54, 2441–2449. [Google Scholar] [CrossRef]
  54. Nordsmark, M.; Bentzen, S.M.; Overgaard, J. Measurement of Human Tumour Oxygenation Status by a Polarographic Needle Electrode. An Analysis of Inter- and Intratumour Heterogeneity. Acta Oncol. 1994, 33, 383–389. [Google Scholar] [CrossRef] [Green Version]
  55. Forster, J.C.; Harriss-Phillips, W.M.; Douglass, M.J.; Bezak, E. A Review of the Development of Tumor Vasculature and Its Effects on the Tumor Microenvironment. Hypoxia 2017, 5, 21–32. [Google Scholar] [CrossRef] [Green Version]
  56. Bayer, C.; Vaupel, P. Acute versus Chronic Hypoxia in Tumors: Controversial Data Concerning Time Frames and Biological Consequences. Strahlenther. Onkol. 2012, 188, 616–627. [Google Scholar] [CrossRef] [PubMed]
  57. Chaplin, D.J.; Durand, R.E.; Olive, P.L. Acute Hypoxia in Tumors: Implications for Modifiers of Radiation Effects. Int. J. Radiat. Oncol. Biol. Phys. 1986, 12, 1279–1282. [Google Scholar] [CrossRef]
  58. Chaplin, D.J.; Olive, P.L.; Durand, R.E. Intermittent Blood Flow in a Murine Tumor: Radiobiological Effects. Cancer Res. 1987, 47, 597–601. [Google Scholar]
  59. Rofstad, E.K.; Gaustad, J.-V.; Egeland, T.A.M.; Mathiesen, B.; Galappathi, K. Tumors Exposed to Acute Cyclic Hypoxic Stress Show Enhanced Angiogenesis, Perfusion and Metastatic Dissemination. Int. J. Cancer 2010, 127, 1535–1546. [Google Scholar] [CrossRef]
  60. Kato, Y.; Yashiro, M.; Fuyuhiro, Y.; Kashiwagi, S.; Matsuoka, J.; Hirakawa, T.; Noda, S.; Aomatsu, N.; Hasegawa, T.; Matsuzaki, T.; et al. Effects of Acute and Chronic Hypoxia on the Radiosensitivity of Gastric and Esophageal Cancer Cells. Anticancer Res. 2011, 31, 3369–3375. [Google Scholar] [PubMed]
  61. Martin, J.D.; Fukumura, D.; Duda, D.G.; Boucher, Y.; Jain, R.K. Reengineering the Tumor Microenvironment to Alleviate Hypoxia and Overcome Cancer Heterogeneity. Cold Spring Harb. Perspect. Med. 2016, 6, a027094. [Google Scholar] [CrossRef] [Green Version]
  62. Brown, J.M.; Giaccia, A.J. The Unique Physiology of Solid Tumors: Opportunities (and Problems) for Cancer Therapy. Cancer Res. 1998, 58, 1408–1416. [Google Scholar]
  63. Semenza, G.L. Hypoxia, Clonal Selection, and the Role of HIF-1 in Tumor Progression. Crit. Rev. Biochem. Mol. Biol. 2000, 35, 71–103. [Google Scholar] [CrossRef] [PubMed]
  64. Semenza, G.L. HIF-1 and Tumor Progression: Pathophysiology and Therapeutics. Trends Mol. Med. 2002, 8, S62–S67. [Google Scholar] [CrossRef]
  65. Zhong, H.; De Marzo, A.M.; Laughner, E.; Lim, M.; Hilton, D.A.; Zagzag, D.; Buechler, P.; Isaacs, W.B.; Semenza, G.L.; Simons, J.W. Overexpression of Hypoxia-Inducible Factor 1alpha in Common Human Cancers and Their Metastases. Cancer Res. 1999, 59, 5830–5835. [Google Scholar] [PubMed]
  66. Talks, K.L.; Turley, H.; Gatter, K.C.; Maxwell, P.H.; Pugh, C.W.; Ratcliffe, P.J.; Harris, A.L. The Expression and Distribution of the Hypoxia-Inducible Factors HIF-1alpha and HIF-2alpha in Normal Human Tissues, Cancers, and Tumor-Associated Macrophages. Am. J. Pathol. 2000, 157, 411–421. [Google Scholar] [CrossRef]
  67. Wang, G.L.; Jiang, B.H.; Rue, E.A.; Semenza, G.L. Hypoxia-Inducible Factor 1 Is a Basic-Helix-Loop-Helix-PAS Heterodimer Regulated by Cellular O2 Tension. Proc. Natl. Acad. Sci. USA 1995, 92, 5510–5514. [Google Scholar] [CrossRef] [Green Version]
  68. Wang, G.L.; Semenza, G.L. Purification and Characterization of Hypoxia-Inducible Factor 1. J. Biol. Chem. 1995, 270, 1230–1237. [Google Scholar] [CrossRef] [Green Version]
  69. Hu, C.-J.; Sataur, A.; Wang, L.; Chen, H.; Simon, M.C. The N-Terminal Transactivation Domain Confers Target Gene Specificity of Hypoxia-Inducible Factors HIF-1alpha and HIF-2alpha. Mol. Biol. Cell 2007, 18, 4528–4542. [Google Scholar] [CrossRef] [Green Version]
  70. Salnikow, K.; Aprelikova, O.; Ivanov, S.; Tackett, S.; Kaczmarek, M.; Karaczyn, A.; Yee, H.; Kasprzak, K.S.; Niederhuber, J. Regulation of Hypoxia-Inducible Genes by ETS1 Transcription Factor. Carcinogenesis 2008, 29, 1493–1499. [Google Scholar] [CrossRef] [Green Version]
  71. Jiang, B.H.; Zheng, J.Z.; Leung, S.W.; Roe, R.; Semenza, G.L. Transactivation and Inhibitory Domains of Hypoxia-Inducible Factor 1alpha. Modulation of Transcriptional Activity by Oxygen Tension. J. Biol. Chem. 1997, 272, 19253–19260. [Google Scholar] [CrossRef] [Green Version]
  72. Dengler, V.L.; Galbraith, M.; Espinosa, J.M. Transcriptional Regulation by Hypoxia Inducible Factors. Crit. Rev. Biochem. Mol. Biol. 2014, 49, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Hu, C.-J.; Wang, L.-Y.; Chodosh, L.A.; Keith, B.; Simon, M.C. Differential Roles of Hypoxia-Inducible Factor 1alpha (HIF-1alpha) and HIF-2alpha in Hypoxic Gene Regulation. Mol. Cell. Biol. 2003, 23, 9361–9374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Takeda, N.; Maemura, K.; Imai, Y.; Harada, T.; Kawanami, D.; Nojiri, T.; Manabe, I.; Nagai, R. Endothelial PAS Domain Protein 1 Gene Promotes Angiogenesis through the Transactivation of Both Vascular Endothelial Growth Factor and Its Receptor, Flt-1. Circ. Res. 2004, 95, 146–153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Koh, M.Y.; Lemos, R., Jr.; Liu, X.; Powis, G. The Hypoxia-Associated Factor Switches Cells from HIF-1α- to HIF-2α-Dependent Signaling Promoting Stem Cell Characteristics, Aggressive Tumor Growth and Invasion. Cancer Res. 2011, 71, 4015–4027. [Google Scholar] [CrossRef] [Green Version]
  76. Koh, M.Y.; Powis, G. Passing the Baton: The HIF Switch. Trends Biochem. Sci. 2012, 37, 364–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Cavadas, M.A.S.; Taylor, C.T.; Cheong, A. Acquisition of Temporal HIF Transcriptional Activity Using a Secreted Luciferase Assay. Methods Mol. Biol. 2018, 1742, 37–44. [Google Scholar] [PubMed]
  78. Gu, Y.Z.; Moran, S.M.; Hogenesch, J.B.; Wartman, L.; Bradfield, C.A. Molecular Characterization and Chromosomal Localization of a Third Alpha-Class Hypoxia Inducible Factor Subunit, HIF3alpha. Gene Expr. 1998, 7, 205–213. [Google Scholar]
  79. Makino, Y.; Cao, R.; Svensson, K.; Bertilsson, G.; Asman, M.; Tanaka, H.; Cao, Y.; Berkenstam, A.; Poellinger, L. Inhibitory PAS Domain Protein Is a Negative Regulator of Hypoxia-Inducible Gene Expression. Nature 2001, 414, 550–554. [Google Scholar] [CrossRef] [PubMed]
  80. Makino, Y.; Kanopka, A.; Wilson, W.J.; 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. [Google Scholar] [CrossRef] [Green Version]
  81. Maynard, M.A.; Evans, A.J.; Hosomi, T.; Hara, S.; Jewett, M.A.S.; Ohh, M. Human HIF-3alpha4 Is a Dominant-Negative Regulator of HIF-1 and Is down-Regulated in Renal Cell Carcinoma. FASEB J. 2005, 19, 1396–1406. [Google Scholar] [CrossRef]
  82. Semenza, G.L.; Wang, G.L. 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. [Google Scholar]
  83. Tian, H.; McKnight, S.L.; Russell, D.W. Endothelial PAS Domain Protein 1 (EPAS1), a Transcription Factor Selectively Expressed in Endothelial Cells. Genes Dev. 1997, 11, 72–82. [Google Scholar] [CrossRef] [Green Version]
  84. Cavadas, M.A.S.; Mesnieres, M.; Crifo, B.; Manresa, M.C.; Selfridge, A.C.; Keogh, C.E.; Fabian, Z.; Scholz, C.C.; Nolan, K.A.; Rocha, L.M.A.; et al. REST Is a Hypoxia-Responsive Transcriptional Repressor. Sci. Rep. 2016, 6, 31355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Wang, Y.; Liu, Y.; Malek, S.N.; Zheng, P.; Liu, Y. Targeting HIF1α Eliminates Cancer Stem Cells in Hematological Malignancies. Cell Stem Cell 2011, 8, 399–411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Mak, P.; Leav, I.; Pursell, B.; Bae, D.; Yang, X.; Taglienti, C.A.; Gouvin, L.M.; Sharma, V.M.; Mercurio, A.M. ERbeta Impedes Prostate Cancer EMT by Destabilizing HIF-1alpha and Inhibiting VEGF-Mediated Snail Nuclear Localization: Implications for Gleason Grading. Cancer Cell 2010, 17, 319–332. [Google Scholar] [CrossRef] [Green Version]
  87. Huang, L.E.; Bindra, R.S.; Glazer, P.M.; Harris, A.L. Hypoxia-Induced Genetic Instability—A Calculated Mechanism Underlying Tumor Progression. J. Mol. Med. 2007, 85, 139–148. [Google Scholar] [CrossRef] [PubMed]
  88. Liao, D.; Corle, C.; Seagroves, T.N.; Johnson, R.S. Hypoxia-Inducible Factor-1alpha Is a Key Regulator of Metastasis in a Transgenic Model of Cancer Initiation and Progression. Cancer Res. 2007, 67, 563–572. [Google Scholar] [CrossRef] [Green Version]
  89. Luo, W.; Hu, H.; Chang, R.; Zhong, J.; Knabel, M.; O’Meally, R.; Cole, R.N.; Pandey, A.; Semenza, G.L. Pyruvate Kinase M2 Is a PHD3-Stimulated Coactivator for Hypoxia-Inducible Factor 1. Cell 2011, 145, 732–744. [Google Scholar] [CrossRef] [Green Version]
  90. Swietach, P.; Vaughan-Jones, R.D.; Harris, A.L. Regulation of Tumor pH and the Role of Carbonic Anhydrase 9. Cancer Metastasis Rev. 2007, 26, 299–310. [Google Scholar] [CrossRef] [PubMed]
  91. Lukashev, D.; Ohta, A.; Sitkovsky, M. Hypoxia-Dependent Anti-Inflammatory Pathways in Protection of Cancerous Tissues. Cancer Metastasis Rev. 2007, 26, 273–279. [Google Scholar] [CrossRef]
  92. Chan, D.A.; Giaccia, A.J. Hypoxia, Gene Expression, and Metastasis. Cancer Metastasis Rev. 2007, 26, 333–339. [Google Scholar] [CrossRef]
  93. Moeller, B.J.; Richardson, R.A.; Dewhirst, M.W. Hypoxia and Radiotherapy: Opportunities for Improved Outcomes in Cancer Treatment. Cancer Metastasis Rev. 2007, 26, 241–248. [Google Scholar] [CrossRef]
  94. Hubbi, M.E.; Kshitiz; Gilkes, D.M.; Rey, S.; Wong, C.C.; Luo, W.; Kim, D.-H.; Dang, C.V.; Levchenko, A.; Semenza, G.L. A Nontranscriptional Role for HIF-1α as a Direct Inhibitor of DNA Replication. Sci. Signal. 2013, 6, ra10. [Google Scholar] [CrossRef] [Green Version]
  95. Huang, L.E.; Gu, J.; Schau, M.; Bunn, H.F. Regulation of Hypoxia-Inducible Factor 1alpha Is Mediated by an O2-Dependent Degradation Domain via the Ubiquitin-Proteasome Pathway. Proc. Natl. Acad. Sci. USA 1998, 95, 7987–7992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. O’Rourke, J.F.; Tian, Y.M.; Ratcliffe, P.J.; Pugh, C.W. Oxygen-Regulated and Transactivating Domains in Endothelial PAS Protein 1: Comparison with Hypoxia-Inducible Factor-1alpha. J. Biol. Chem. 1999, 274, 2060–2071. [Google Scholar] [CrossRef] [Green Version]
  97. Ivan, M.; Kondo, K.; Yang, H.; Kim, W.; Valiando, J.; Ohh, M.; Salic, A.; Asara, J.M.; Lane, W.S.; Kaelin, W.G., Jr. HIFalpha Targeted for VHL-Mediated Destruction by Proline Hydroxylation: Implications for O2 Sensing. Science 2001, 292, 464–468. [Google Scholar] [CrossRef] [PubMed]
  98. Jaakkola, P.; Mole, D.R.; Tian, Y.M.; Wilson, M.I.; Gielbert, J.; Gaskell, S.J.; von Kriegsheim, A.; Hebestreit, H.F.; Mukherji, M.; Schofield, C.J.; et al. Targeting of HIF-Alpha to the von Hippel-Lindau Ubiquitylation Complex by O2-Regulated Prolyl Hydroxylation. Science 2001, 292, 468–472. [Google Scholar] [CrossRef] [PubMed]
  99. Mahon, P.C.; Hirota, K.; Semenza, G.L. FIH-1: A Novel Protein That Interacts with HIF-1alpha and VHL to Mediate Repression of HIF-1 Transcriptional Activity. Genes Dev. 2001, 15, 2675–2686. [Google Scholar] [CrossRef] [Green Version]
  100. Wang, G.L.; Semenza, G.L. General Involvement of Hypoxia-Inducible Factor 1 in Transcriptional Response to Hypoxia. Proc. Natl. Acad. Sci. USA 1993, 90, 4304–4308. [Google Scholar] [CrossRef] [Green Version]
  101. Wang, G.L.; Semenza, G.L. Characterization of Hypoxia-Inducible Factor 1 and Regulation of DNA Binding Activity by Hypoxia. J. Biol. Chem. 1993, 268, 21513–21518. [Google Scholar] [CrossRef]
  102. Ema, M.; Taya, S.; Yokotani, N.; Sogawa, K.; Matsuda, Y.; Fujii-Kuriyama, Y. A Novel bHLH-PAS Factor with Close Sequence Similarity to Hypoxia-Inducible Factor 1alpha Regulates the VEGF Expression and Is Potentially Involved in Lung and Vascular Development. Proc. Natl. Acad. Sci. USA 1997, 94, 4273–4278. [Google Scholar] [CrossRef] [Green Version]
  103. Flamme, I.; Fröhlich, T.; von Reutern, M.; Kappel, A.; Damert, A.; Risau, W. HRF, a Putative Basic Helix-Loop-Helix-PAS-Domain Transcription Factor Is Closely Related to Hypoxia-Inducible Factor-1 Alpha and Developmentally Expressed in Blood Vessels. Mech. Dev. 1997, 63, 51–60. [Google Scholar] [CrossRef]
  104. Schito, L.; Semenza, G.L. Hypoxia-Inducible Factors: Master Regulators of Cancer Progression. Trends Cancer Res. 2016, 2, 758–770. [Google Scholar] [CrossRef] [Green Version]
  105. Goldberg, M.A.; Dunning, S.P.; Bunn, H.F. Regulation of the Erythropoietin Gene: Evidence That the Oxygen Sensor Is a Heme Protein. Science 1988, 242, 1412–1415. [Google Scholar] [CrossRef]
  106. Salnikow, K.; An, W.G.; Melillo, G.; Blagosklonny, M.V.; Costa, M. Nickel-Induced Transformation Shifts the Balance between HIF-1 and p53 Transcription Factors. Carcinogenesis 1999, 20, 1819–1823. [Google Scholar] [CrossRef] [Green Version]
  107. Befani, C.; Mylonis, I.; Gkotinakou, I.-M.; Georgoulias, P.; Hu, C.-J.; Simos, G.; Liakos, P. Cobalt Stimulates HIF-1-Dependent but Inhibits HIF-2-Dependent Gene Expression in Liver Cancer Cells. Int. J. Biochem. Cell Biol. 2013, 45, 2359–2368. [Google Scholar] [CrossRef] [Green Version]
  108. Kaczmarek, M.; Timofeeva, O.A.; Karaczyn, A.; Malyguine, A.; Kasprzak, K.S.; Salnikow, K. The Role of Ascorbate in the Modulation of HIF-1alpha Protein and HIF-Dependent Transcription by chromium(VI) and nickel(II). Free Radic. Biol. Med. 2007, 42, 1246–1257. [Google Scholar] [CrossRef] [Green Version]
  109. Li, Q.; Chen, H.; Huang, X.; Costa, M. Effects of 12 Metal Ions on Iron Regulatory Protein 1 (IRP-1) and Hypoxia-Inducible Factor-1 Alpha (HIF-1alpha) and HIF-Regulated Genes. Toxicol. Appl. Pharmacol. 2006, 213, 245–255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Salnikow, K.; Donald, S.P.; Bruick, R.K.; Zhitkovich, A.; Phang, J.M.; Kasprzak, K.S. Depletion of Intracellular Ascorbate by the Carcinogenic Metals Nickel and Cobalt Results in the Induction of Hypoxic Stress. J. Biol. Chem. 2004, 279, 40337–40344. [Google Scholar] [CrossRef] [Green Version]
  111. Kaczmarek, M.; Cachau, R.E.; Topol, I.A.; Kasprzak, K.S.; Ghio, A.; Salnikow, K. Metal Ions-Stimulated Iron Oxidation in Hydroxylases Facilitates Stabilization of HIF-1 Alpha Protein. Toxicol. Sci. 2009, 107, 394–403. [Google Scholar] [CrossRef] [Green Version]
  112. Chandel, N.S.; McClintock, D.S.; Feliciano, C.E.; Wood, T.M.; Melendez, J.A.; Rodriguez, A.M.; Schumacker, P.T. Reactive Oxygen Species Generated at Mitochondrial Complex III Stabilize Hypoxia-Inducible Factor-1alpha during Hypoxia: A Mechanism of O2 Sensing. J. Biol. Chem. 2000, 275, 25130–25138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Guzy, R.D.; Hoyos, B.; Robin, E.; Chen, H.; Liu, L.; Mansfield, K.D.; Simon, M.C.; Hammerling, U.; Schumacker, P.T. Mitochondrial Complex III Is Required for Hypoxia-Induced ROS Production and Cellular Oxygen Sensing. Cell Metab. 2005, 1, 401–408. [Google Scholar] [CrossRef] [Green Version]
  114. Kaewpila, S.; Venkataraman, S.; Buettner, G.R.; Oberley, L.W. Manganese Superoxide Dismutase Modulates Hypoxia-Inducible Factor-1 Alpha Induction via Superoxide. Cancer Res. 2008, 68, 2781–2788. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Chua, Y.L.; Dufour, E.; Dassa, E.P.; Rustin, P.; Jacobs, H.T.; Taylor, C.T.; Hagen, T. Stabilization of Hypoxia-Inducible Factor-1alpha Protein in Hypoxia Occurs Independently of Mitochondrial Reactive Oxygen Species Production. J. Biol. Chem. 2010, 285, 31277–31284. [Google Scholar] [CrossRef] [Green Version]
  116. Bell, E.L.; Klimova, T.A.; Eisenbart, J.; Moraes, C.T.; Murphy, M.P.; Budinger, G.R.S.; Chandel, N.S. The Qo Site of the Mitochondrial Complex III Is Required for the Transduction of Hypoxic Signaling via Reactive Oxygen Species Production. J. Cell Biol. 2007, 177, 1029–1036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Sasabe, E.; Yang, Z.; Ohno, S.; Yamamoto, T. Reactive Oxygen Species Produced by the Knockdown of Manganese-Superoxide Dismutase up-Regulate Hypoxia-Inducible Factor-1alpha Expression in Oral Squamous Cell Carcinoma Cells. Free Radic. Biol. Med. 2010, 48, 1321–1329. [Google Scholar] [CrossRef] [PubMed]
  118. Masson, N.; Singleton, R.S.; Sekirnik, R.; Trudgian, D.C.; Ambrose, L.J.; Miranda, M.X.; Tian, Y.-M.; Kessler, B.M.; Schofield, C.J.; Ratcliffe, P.J. The FIH Hydroxylase Is a Cellular Peroxide Sensor That Modulates HIF Transcriptional Activity. EMBO Rep. 2012, 13, 251–257. [Google Scholar] [CrossRef] [PubMed]
  119. Sandau, K.B.; Fandrey, J.; Brüne, B. Accumulation of HIF-1alpha under the Influence of Nitric Oxide. Blood 2001, 97, 1009–1015. [Google Scholar] [CrossRef] [PubMed]
  120. Palmer, L.A.; Gaston, B.; Johns, R.A. Normoxic Stabilization of Hypoxia-Inducible Factor-1 Expression and Activity: Redox-Dependent Effect of Nitrogen Oxides. Mol. Pharmacol. 2000, 58, 1197–1203. [Google Scholar] [CrossRef]
  121. Sumbayev, V.V.; Budde, A.; Zhou, J.; Brüne, B. HIF-1 Alpha Protein as a Target for S-Nitrosation. FEBS Lett. 2003, 535, 106–112. [Google Scholar] [CrossRef] [Green Version]
  122. Metzen, E.; Zhou, J.; Jelkmann, W.; Fandrey, J.; Brüne, B. Nitric Oxide Impairs Normoxic Degradation of HIF-1alpha by Inhibition of Prolyl Hydroxylases. Mol. Biol. Cell 2003, 14, 3470–3481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Olson, N.; van der Vliet, A. Interactions between Nitric Oxide and Hypoxia-Inducible Factor Signaling Pathways in Inflammatory Disease. Nitric Oxide 2011, 25, 125–137. [Google Scholar] [CrossRef] [Green Version]
  124. Movafagh, S.; Crook, S.; Vo, K. Regulation of Hypoxia-Inducible Factor-1a by Reactive Oxygen Species: New Developments in an Old Debate. J. Cell. Biochem. 2015, 116, 696–703. [Google Scholar] [CrossRef]
  125. Li, J.; Brown, L.F.; Hibberd, M.G.; Grossman, J.D.; Morgan, J.P.; Simons, M. VEGF, Flk-1, and Flt-1 Expression in a Rat Myocardial Infarction Model of Angiogenesis. Am. J. Physiol. 1996, 270, H1803–H1811. [Google Scholar] [CrossRef]
  126. Kim, C.-H.; Cho, Y.-S.; Chun, Y.-S.; Park, J.-W.; Kim, M.-S. Early Expression of Myocardial HIF-1alpha in Response to Mechanical Stresses: Regulation by Stretch-Activated Channels and the Phosphatidylinositol 3-Kinase Signaling Pathway. Circ. Res. 2002, 90, E25–E33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Chang, H.; Shyu, K.-G.; Wang, B.-W.; Kuan, P. Regulation of Hypoxia-Inducible Factor-1alpha by Cyclical Mechanical Stretch in Rat Vascular Smooth Muscle Cells. Clin. Sci. 2003, 105, 447–456. [Google Scholar] [CrossRef] [Green Version]
  128. Milkiewicz, M.; Doyle, J.L.; Fudalewski, T.; Ispanovic, E.; Aghasi, M.; Haas, T.L. HIF-1alpha and HIF-2alpha Play a Central Role in Stretch-Induced but Not Shear-Stress-Induced Angiogenesis in Rat Skeletal Muscle. J. Physiol. 2007, 583, 753–766. [Google Scholar] [CrossRef] [PubMed]
  129. Deroo, B.J.; Korach, K.S. Estrogen Receptors and Human Disease. J. Clin. Investig. 2006, 116, 561–570. [Google Scholar] [CrossRef] [Green Version]
  130. Seton-Rogers, S. Breast Cancer: Untangling the Role of Progesterone Receptors. Nat. Rev. Cancer 2015, 15, 456. [Google Scholar] [CrossRef]
  131. D’Uva, G.; Lauriola, M. Towards the Emerging Crosstalk: ERBB Family and Steroid Hormones. Semin. Cell Dev. Biol. 2016, 50, 143–152. [Google Scholar] [CrossRef] [PubMed]
  132. Lappano, R.; De Marco, P.; De Francesco, E.M.; Chimento, A.; Pezzi, V.; Maggiolini, M. Cross-Talk between GPER and Growth Factor Signaling. J. Steroid Biochem. Mol. Biol. 2013, 137, 50–56. [Google Scholar] [CrossRef]
  133. Hawsawi, Y.; El-Gendy, R.; Twelves, C.; Speirs, V.; Beattie, J. Insulin-like Growth Factor—Oestradiol Crosstalk and Mammary Gland Tumourigenesis. Biochim. Biophys. Acta 2013, 1836, 345–353. [Google Scholar] [CrossRef]
  134. Bartella, V.; De Marco, P.; Malaguarnera, R.; Belfiore, A.; Maggiolini, M. New Advances on the Functional Cross-Talk between Insulin-like Growth Factor-I and Estrogen Signaling in Cancer. Cell. Signal. 2012, 24, 1515–1521. [Google Scholar] [CrossRef]
  135. Liang, J.; Shang, Y. Estrogen and Cancer. Annu. Rev. Physiol. 2013, 75, 225–240. [Google Scholar] [CrossRef] [Green Version]
  136. Chen, G.G.; Zeng, Q.; Tse, G.M. Estrogen and Its Receptors in Cancer. Med. Res. Rev. 2008, 28, 954–974. [Google Scholar] [CrossRef]
  137. Heldring, N.; Pike, A.; Andersson, S.; Matthews, J.; Cheng, G.; Hartman, J.; Tujague, M.; Ström, A.; Treuter, E.; Warner, M.; et al. Estrogen Receptors: How Do They Signal and What Are Their Targets. Physiol. Rev. 2007, 87, 905–931. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Björnström, L.; Sjöberg, M. Mechanisms of Estrogen Receptor Signaling: Convergence of Genomic and Nongenomic Actions on Target Genes. Mol. Endocrinol. 2005, 19, 833–842. [Google Scholar] [CrossRef] [Green Version]
  139. Hammes, S.R.; Levin, E.R. Minireview: Recent Advances in Extranuclear Steroid Receptor Actions. Endocrinology 2011, 152, 4489–4495. [Google Scholar] [CrossRef] [Green Version]
  140. Prossnitz, E.R.; Barton, M. The G-Protein-Coupled Estrogen Receptor GPER in Health and Disease. Nat. Rev. Endocrinol. 2011, 7, 715–726. [Google Scholar] [CrossRef] [Green Version]
  141. Barton, M.; Filardo, E.J.; Lolait, S.J.; Thomas, P.; Maggiolini, M.; Prossnitz, E.R. Twenty Years of the G Protein-Coupled Estrogen Receptor GPER: Historical and Personal Perspectives. J. Steroid Biochem. Mol. Biol. 2018, 176, 4–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Musgrove, E.A.; Caldon, C.E.; Barraclough, J.; Stone, A.; Sutherland, R.L. Cyclin D as a Therapeutic Target in Cancer. Nat. Rev. Cancer 2011, 11, 558–572. [Google Scholar] [CrossRef]
  143. Caldon, C.E.; Sutherland, R.L.; Musgrove, E. Cell Cycle Proteins in Epithelial Cell Differentiation: Implications for Breast Cancer. Cell Cycle 2010, 9, 1918–1928. [Google Scholar] [CrossRef]
  144. Gompel, A.; Somaï, S.; Chaouat, M.; Kazem, A.; Kloosterboer, H.J.; Beusman, I.; Forgez, P.; Mimoun, M.; Rostène, W. Hormonal Regulation of Apoptosis in Breast Cells and Tissues. Steroids 2000, 65, 593–598. [Google Scholar] [CrossRef]
  145. Miller, V.M.; Duckles, S.P. Vascular Actions of Estrogens: Functional Implications. Pharmacol. Rev. 2008, 60, 210–241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Bendrik, C.; Dabrosin, C. Estradiol Increases IL-8 Secretion of Normal Human Breast Tissue and Breast Cancer In Vivo. J. Immunol. 2009, 182, 371–378. [Google Scholar] [CrossRef] [Green Version]
  147. Lindahl, G.; Saarinen, N.; Abrahamsson, A.; Dabrosin, C. Tamoxifen, Flaxseed, and the Lignan Enterolactone Increase Stroma- and Cancer Cell-Derived IL-1Ra and Decrease Tumor Angiogenesis in Estrogen-Dependent Breast Cancer. Cancer Res. 2011, 71, 51–60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Wang, X.; Belguise, K.; Kersual, N.; Kirsch, K.H.; Mineva, N.D.; Galtier, F.; Chalbos, D.; Sonenshein, G.E. Oestrogen Signalling Inhibits Invasive Phenotype by Repressing RelB and Its Target BCL2. Nat. Cell Biol. 2007, 9, 470–478. [Google Scholar] [CrossRef] [Green Version]
  149. McCune, K.; Mehta, R.; Thorat, M.A.; Badve, S.; Nakshatri, H. Loss of ERα and FOXA1 Expression in a Progression Model of Luminal Type Breast Cancer: Insights from PyMT Transgenic Mouse Model. Oncol. Rep. 2010, 24, 1233–1239. [Google Scholar]
  150. Guttilla, I.K.; Adams, B.D.; White, B.A. ERα, microRNAs, and the Epithelial-Mesenchymal Transition in Breast Cancer. Trends Endocrinol. Metab. 2012, 23, 73–82. [Google Scholar] [CrossRef] [PubMed]
  151. Pandey, D.P.; Lappano, R.; Albanito, L.; Madeo, A.; Maggiolini, M.; Picard, D. Estrogenic GPR30 Signalling Induces Proliferation and Migration of Breast Cancer Cells through CTGF. EMBO J. 2009, 28, 523–532. [Google Scholar] [CrossRef] [Green Version]
  152. Filardo, E.J. A Role for G-Protein Coupled Estrogen Receptor (GPER) in Estrogen-Induced Carcinogenesis: Dysregulated Glandular Homeostasis, Survival and Metastasis. J. Steroid Biochem. Mol. Biol. 2018, 176, 38–48. [Google Scholar] [CrossRef]
  153. Lappano, R.; Maggiolini, M. GPER Is Involved in the Functional Liaison between Breast Tumor Cells and Cancer-Associated Fibroblasts (CAFs). J. Steroid Biochem. Mol. Biol. 2018, 176, 49–56. [Google Scholar] [CrossRef]
  154. Lappano, R.; Talia, M.; Cirillo, F.; Rigiracciolo, D.C.; Scordamaglia, D.; Guzzi, R.; Miglietta, A.M.; De Francesco, E.M.; Belfiore, A.; Sims, A.H.; et al. The IL1β-IL1R Signaling Is Involved in the Stimulatory Effects Triggered by Hypoxia in Breast Cancer Cells and Cancer-Associated Fibroblasts (CAFs). J. Exp. Clin. Cancer Res. 2020, 39, 153. [Google Scholar] [CrossRef]
  155. De Francesco, E.M.; Pellegrino, M.; Santolla, M.F.; Lappano, R.; Ricchio, E.; Abonante, S.; Maggiolini, M. GPER Mediates Activation of HIF1α/VEGF Signaling by Estrogens. Cancer Res. 2014, 74, 4053–4064. [Google Scholar] [CrossRef] [Green Version]
  156. De Francesco, E.M.; Lappano, R.; Santolla, M.F.; Marsico, S.; Caruso, A.; Maggiolini, M. HIF-1α/GPER Signaling Mediates the Expression of VEGF Induced by Hypoxia in Breast Cancer Associated Fibroblasts (CAFs). Breast Cancer Res. 2013, 15, R64. [Google Scholar] [CrossRef] [Green Version]
  157. De Marco, P.; Lappano, R.; De Francesco, E.M.; Cirillo, F.; Pupo, M.; Avino, S.; Vivacqua, A.; Abonante, S.; Picard, D.; Maggiolini, M. GPER Signalling in Both Cancer-Associated Fibroblasts and Breast Cancer Cells Mediates a Feedforward IL1β/IL1R1 Response. Sci. Rep. 2016, 6, 24354. [Google Scholar] [CrossRef] [Green Version]
  158. Egloff, A.M.; Rothstein, M.E.; Seethala, R.; Siegfried, J.M.; Grandis, J.R.; Stabile, L.P. Cross-Talk between Estrogen Receptor and Epidermal Growth Factor Receptor in Head and Neck Squamous Cell Carcinoma. Clin. Cancer Res. 2009, 15, 6529–6540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Lanzino, M.; Morelli, C.; Garofalo, C.; Panno, M.L.; Mauro, L.; Andò, S.; Sisci, D. Interaction between Estrogen Receptor Alpha and insulin/IGF Signaling in Breast Cancer. Curr. Cancer Drug Targets 2008, 8, 597–610. [Google Scholar] [CrossRef] [PubMed]
  160. Siegfried, J.M.; Farooqui, M.; Rothenberger, N.J.; Dacic, S.; Stabile, L.P. Interaction between the Estrogen Receptor and Fibroblast Growth Factor Receptor Pathways in Non-Small Cell Lung Cancer. Oncotarget 2017, 8, 24063–24076. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Santolla, M.F.; Vivacqua, A.; Lappano, R.; Rigiracciolo, D.C.; Cirillo, F.; Galli, G.R.; Talia, M.; Brunetti, G.; Miglietta, A.M.; Belfiore, A.; et al. GPER Mediates a Feedforward FGF2/FGFR1 Paracrine Activation Coupling CAFs to Cancer Cells toward Breast Tumor Progression. Cells 2019, 8, 223. [Google Scholar] [CrossRef] [Green Version]
  162. De Marco, P.; Bartella, V.; Vivacqua, A.; Lappano, R.; Santolla, M.F.; Morcavallo, A.; Pezzi, V.; Belfiore, A.; Maggiolini, M. Insulin-like Growth Factor-I Regulates GPER Expression and Function in Cancer Cells. Oncogene 2013, 32, 678–688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Salomon, D.S.; Brandt, R.; Ciardiello, F.; Normanno, N. Epidermal Growth Factor-Related Peptides and Their Receptors in Human Malignancies. Crit. Rev. Oncol. Hematol. 1995, 19, 183–232. [Google Scholar] [CrossRef]
  164. Gullick, W.J. The Epidermal Growth Factor System of Ligands and Receptors in Cancer. Eur. J. Cancer 2009, 45 (Suppl. S1), 205–210. [Google Scholar] [CrossRef]
  165. Belfiore, A.; Frasca, F.; Pandini, G.; Sciacca, L.; Vigneri, R. Insulin Receptor Isoforms and Insulin Receptor/insulin-like Growth Factor Receptor Hybrids in Physiology and Disease. Endocr. Rev. 2009, 30, 586–623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Pollak, M. Insulin and Insulin-like Growth Factor Signalling in Neoplasia. Nat. Rev. Cancer 2008, 8, 915–928. [Google Scholar] [CrossRef] [PubMed]
  167. Turner, N.; Grose, R. Fibroblast Growth Factor Signalling: From Development to Cancer. Nat. Rev. Cancer 2010, 10, 116–129. [Google Scholar] [CrossRef] [PubMed]
  168. Lindsey, S.; Langhans, S.A. Epidermal Growth Factor Signaling in Transformed Cells. Int. Rev. Cell Mol. Biol. 2015, 314, 1–41. [Google Scholar] [PubMed] [Green Version]
  169. Krook, M.A.; Reeser, J.W.; Ernst, G.; Barker, H.; Wilberding, M.; Li, G.; Chen, H.-Z.; Roychowdhury, S. Fibroblast Growth Factor Receptors in Cancer: Genetic Alterations, Diagnostics, Therapeutic Targets and Mechanisms of Resistance. Br. J. Cancer 2021, 124, 880–892. [Google Scholar] [CrossRef]
  170. Heldin, C.-H.; Lennartsson, J.; Westermark, B. Involvement of Platelet-Derived Growth Factor Ligands and Receptors in Tumorigenesis. J. Intern. Med. 2018, 283, 16–44. [Google Scholar] [CrossRef] [Green Version]
  171. Franovic, A.; Gunaratnam, L.; Smith, K.; Robert, I.; Patten, D.; Lee, S. Translational up-Regulation of the EGFR by Tumor Hypoxia Provides a Nonmutational Explanation for Its Overexpression in Human Cancer. Proc. Natl. Acad. Sci. USA 2007, 104, 13092–13097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Smith, K.; Gunaratnam, L.; Morley, M.; Franovic, A.; Mekhail, K.; Lee, S. Silencing of Epidermal Growth Factor Receptor Suppresses Hypoxia-Inducible Factor-2-Driven VHL-/- Renal Cancer. Cancer Res. 2005, 65, 5221–5230. [Google Scholar] [CrossRef] [Green Version]
  173. Franovic, A.; Holterman, C.E.; Payette, J.; Lee, S. Human Cancers Converge at the HIF-2alpha Oncogenic Axis. Proc. Natl. Acad. Sci. USA 2009, 106, 21306–21311. [Google Scholar] [CrossRef] [Green Version]
  174. Laughner, E.; Taghavi, P.; Chiles, K.; Mahon, P.C.; Semenza, G.L. 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. [Google Scholar] [CrossRef] [Green Version]
  175. Fukuda, R.; Hirota, K.; Fan, F.; Jung, Y.D.; Ellis, L.M.; Semenza, G.L. Insulin-like Growth Factor 1 Induces Hypoxia-Inducible Factor 1-Mediated Vascular Endothelial Growth Factor Expression, Which Is Dependent on MAP Kinase and Phosphatidylinositol 3-Kinase Signaling in Colon Cancer Cells. J. Biol. Chem. 2002, 277, 38205–38211. [Google Scholar] [CrossRef] [Green Version]
  176. Shi, Y.-H.; Bingle, L.; Gong, L.-H.; Wang, Y.-X.; Corke, K.P.; Fang, W.-G. Basic FGF Augments Hypoxia Induced HIF-1-Alpha Expression and VEGF Release in T47D Breast Cancer Cells. Pathology 2007, 39, 396–400. [Google Scholar] [CrossRef]
  177. Bos, R.; van Diest, P.J.; de Jong, J.S.; van der Groep, P.; van der Valk, P.; van der Wall, E. Hypoxia-Inducible Factor-1alpha Is Associated with Angiogenesis, and Expression of bFGF, PDGF-BB, and EGFR in Invasive Breast Cancer. Histopathology 2005, 46, 31–36. [Google Scholar] [CrossRef] [PubMed]
  178. Hirami, Y.; Aoe, M.; Tsukuda, K.; Hara, F.; Otani, Y.; Koshimune, R.; Hanabata, T.; Nagahiro, I.; Sano, Y.; Date, H.; et al. Relation of Epidermal Growth Factor Receptor, Phosphorylated-Akt, and Hypoxia-Inducible Factor-1alpha in Non-Small Cell Lung Cancers. Cancer Lett. 2004, 214, 157–164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Mabjeesh, N.J.; Willard, M.T.; Frederickson, C.E.; Zhong, H.; Simons, J.W. Androgens Stimulate Hypoxia-Inducible Factor 1 Activation via Autocrine Loop of Tyrosine Kinase Receptor/phosphatidylinositol 3′-Kinase/protein Kinase B in Prostate Cancer Cells. Clin. Cancer Res. 2003, 9, 2416–2425. [Google Scholar] [PubMed]
  180. Hua, K.; Din, J.; Cao, Q.; Feng, W.; Zhang, Y.; Yao, L.; Huang, Y.; Zhao, Y.; Feng, Y. Estrogen and Progestin Regulate HIF-1alpha Expression in Ovarian Cancer Cell Lines via the Activation of Akt Signaling Transduction Pathway. Oncol. Rep. 2009, 21, 893–898. [Google Scholar] [PubMed] [Green Version]
  181. Von Wahlde, M.-K.; Hülsewig, C.; Ruckert, C.; Götte, M.; Kiesel, L.; Bernemann, C. The Anti-Androgen Drug Dutasteride Renders Triple Negative Breast Cancer Cells More Sensitive to Chemotherapy via Inhibition of HIF-1α-/VEGF-Signaling. Gynecol. Endocrinol. 2015, 31, 160–164. [Google Scholar] [CrossRef] [PubMed]
  182. Ragnum, H.B.; Røe, K.; Holm, R.; Vlatkovic, L.; Nesland, J.M.; Aarnes, E.-K.; Ree, A.H.; Flatmark, K.; Seierstad, T.; Lilleby, W.; et al. Hypoxia-Independent Downregulation of Hypoxia-Inducible Factor 1 Targets by Androgen Deprivation Therapy in Prostate Cancer. Int. J. Radiat. Oncol. Biol. Phys. 2013, 87, 753–760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Rajoria, S.; Hanly, E.; Nicolini, A.; George, A.L.; Geliebter, J.; Shin, E.J.; Suriano, R.; Carpi, A.; Tiwari, R.K. Interlinking of Hypoxia and Estrogen in Thyroid Cancer Progression. Curr. Med. Chem. 2014, 21, 1351–1360. [Google Scholar] [CrossRef]
  184. George, A.L.; Rajoria, S.; Suriano, R.; Mittleman, A.; Tiwari, R.K. Hypoxia and Estrogen Are Functionally Equivalent in Breast Cancer-Endothelial Cell Interdependence. Mol. Cancer 2012, 11, 80. [Google Scholar] [CrossRef] [Green Version]
  185. Dewangan, J.; Kaushik, S.; Rath, S.K.; Balapure, A.K. Centchroman Regulates Breast Cancer Angiogenesis via Inhibition of HIF-1α/VEGFR2 Signalling Axis. Life Sci. 2018, 193, 9–19. [Google Scholar] [CrossRef]
  186. Fuady, J.H.; Gutsche, K.; Santambrogio, S.; Varga, Z.; Hoogewijs, D.; Wenger, R.H. Estrogen-Dependent Downregulation of Hypoxia-Inducible Factor (HIF)-2α in Invasive Breast Cancer Cells. Oncotarget 2016, 7, 31153–31165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Stiehl, D.P.; Bordoli, M.R.; Abreu-Rodríguez, I.; Wollenick, K.; Schraml, P.; Gradin, K.; Poellinger, L.; Kristiansen, G.; Wenger, R.H. Non-Canonical HIF-2α Function Drives Autonomous Breast Cancer Cell Growth via an AREG-EGFR/ErbB4 Autocrine Loop. Oncogene 2012, 31, 2283–2297. [Google Scholar] [CrossRef] [Green Version]
  188. Seifeddine, R.; Dreiem, A.; Tomkiewicz, C.; Fulchignoni-Lataud, M.-C.; Brito, I.; Danan, J.-L.; Favaudon, V.; Barouki, R.; Massaad-Massade, L. Hypoxia and Estrogen Co-Operate to Regulate Gene Expression in T-47D Human Breast Cancer Cells. J. Steroid Biochem. Mol. Biol. 2007, 104, 169–179. [Google Scholar] [CrossRef] [PubMed]
  189. Yang, J.; Jubb, A.M.; Pike, L.; Buffa, F.M.; Turley, H.; Baban, D.; Leek, R.; Gatter, K.C.; Ragoussis, J.; Harris, A.L. The Histone Demethylase JMJD2B Is Regulated by Estrogen Receptor Alpha and Hypoxia, and Is a Key Mediator of Estrogen Induced Growth. Cancer Res. 2010, 70, 6456–6466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Yang, J.; Harris, A.L.; Davidoff, A.M. Hypoxia and Hormone-Mediated Pathways Converge at the Histone Demethylase KDM4B in Cancer. Int. J. Mol. Sci. 2018, 19, 240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Kawazu, M.; Saso, K.; Tong, K.I.; McQuire, T.; Goto, K.; Son, D.-O.; Wakeham, A.; Miyagishi, M.; Mak, T.W.; Okada, H. Histone Demethylase JMJD2B Functions as a Co-Factor of Estrogen Receptor in Breast Cancer Proliferation and Mammary Gland Development. PLoS ONE 2011, 6, e17830. [Google Scholar] [CrossRef]
  192. Yang, J.; AlTahan, A.; Jones, D.T.; Buffa, F.M.; Bridges, E.; Interiano, R.B.; Qu, C.; Vogt, N.; Li, J.-L.; Baban, D.; et al. Estrogen Receptor-α Directly Regulates the Hypoxia-Inducible Factor 1 Pathway Associated with Antiestrogen Response in Breast Cancer. Proc. Natl. Acad. Sci. USA 2015, 112, 15172–15177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Generali, D.; Buffa, F.M.; Berruti, A.; Brizzi, M.P.; Campo, L.; Bonardi, S.; Bersiga, A.; Allevi, G.; Milani, M.; Aguggini, S.; et al. 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. [Google Scholar] [CrossRef]
  194. Generali, D.; Berruti, A.; Brizzi, M.P.; Campo, L.; Bonardi, S.; Wigfield, S.; Bersiga, A.; Allevi, G.; Milani, M.; Aguggini, S.; et al. 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. [Google Scholar] [CrossRef] [Green Version]
  195. Bos, R.; Zhong, H.; Hanrahan, C.F.; Mommers, E.C.; Semenza, G.L.; Pinedo, H.M.; Abeloff, M.D.; Simons, J.W.; van Diest, P.J.; van der Wall, E. Levels of Hypoxia-Inducible Factor-1 Alpha during Breast Carcinogenesis. J. Natl. Cancer Inst. 2001, 93, 309–314. [Google Scholar] [CrossRef] [PubMed]
  196. Bialesova, L.; Xu, L.; Gustafsson, J.-Å.; Haldosen, L.-A.; Zhao, C.; Dahlman-Wright, K. Estrogen Receptor β2 Induces Proliferation and Invasiveness of Triple Negative Breast Cancer Cells: Association with Regulation of PHD3 and HIF-1α. Oncotarget 2017, 8, 76622–76633. [Google Scholar] [CrossRef] [Green Version]
  197. Dey, P.; Velazquez-Villegas, L.A.; Faria, M.; Turner, A.; Jonsson, P.; Webb, P.; Williams, C.; Gustafsson, J.-Å.; Ström, A.M. Estrogen Receptor β2 Induces Hypoxia Signature of Gene Expression by Stabilizing HIF-1α in Prostate Cancer. PLoS ONE 2015, 10, e0128239. [Google Scholar]
  198. Zou, C.; Yu, S.; Xu, Z.; Wu, D.; Ng, C.-F.; Yao, X.; Yew, D.T.; Vanacker, J.-M.; Chan, F.L. ERRα Augments HIF-1 Signalling by Directly Interacting with HIF-1α in Normoxic and Hypoxic Prostate Cancer Cells. J. Pathol. 2014, 233, 61–73. [Google Scholar] [CrossRef]
  199. Zhong, H.; Chiles, K.; Feldser, D.; Laughner, E.; Hanrahan, C.; Georgescu, M.M.; Simons, J.W.; Semenza, G.L. 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. [Google Scholar]
  200. Phillips, R.J.; Mestas, J.; Gharaee-Kermani, M.; Burdick, M.D.; Sica, A.; Belperio, J.A.; Keane, M.P.; Strieter, R.M. Epidermal Growth Factor and Hypoxia-Induced Expression of CXC Chemokine Receptor 4 on Non-Small Cell Lung Cancer Cells Is Regulated by the Phosphatidylinositol 3-kinase/PTEN/AKT/mammalian Target of Rapamycin Signaling Pathway and Activation of Hypoxia Inducible Factor-1alpha. J. Biol. Chem. 2005, 280, 22473–22481. [Google Scholar]
  201. Han, Z.-B.; Ren, H.; Zhao, H.; Chi, Y.; Chen, K.; Zhou, B.; Liu, Y.-J.; Zhang, L.; Xu, B.; Liu, B.; et al. Hypoxia-Inducible Factor (HIF)-1 Alpha Directly Enhances the Transcriptional Activity of Stem Cell Factor (SCF) in Response to Hypoxia and Epidermal Growth Factor (EGF). Carcinogenesis 2008, 29, 1853–1861. [Google Scholar] [CrossRef]
  202. Zhao, F.-L.; Qin, C.-F. EGF Promotes HIF-1α Expression in Colorectal Cancer Cells and Tumor Metastasis by Regulating Phosphorylation of STAT3. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 1055–1062. [Google Scholar] [PubMed]
  203. Peng, X.-H.; Karna, P.; Cao, Z.; Jiang, B.-H.; Zhou, M.; Yang, L. Cross-Talk between Epidermal Growth Factor Receptor and Hypoxia-Inducible Factor-1alpha Signal Pathways Increases Resistance to Apoptosis by up-Regulating Survivin Gene Expression. J. Biol. Chem. 2006, 281, 25903–25914. [Google Scholar] [CrossRef] [Green Version]
  204. Swinson, D.E.B.; O’Byrne, K.J. Interactions between Hypoxia and Epidermal Growth Factor Receptor in Non-Small-Cell Lung Cancer. Clin. Lung Cancer 2006, 7, 250–256. [Google Scholar] [CrossRef]
  205. Hu, W.; Zheng, S.; Guo, H.; Dai, B.; Ni, J.; Shi, Y.; Bian, H.; Li, L.; Shen, Y.; Wu, M.; et al. PLAGL2-EGFR-HIF-1/2α Signaling Loop Promotes HCC Progression and Erlotinib Insensitivity. Hepatology 2021, 73, 674–691. [Google Scholar] [CrossRef]
  206. Wang, Y.; Roche, O.; Xu, C.; Moriyama, E.H.; Heir, P.; Chung, J.; Roos, F.C.; Chen, Y.; Finak, G.; Milosevic, M.; et al. Hypoxia Promotes Ligand-Independent EGF Receptor Signaling via Hypoxia-Inducible Factor-Mediated Upregulation of Caveolin-1. Proc. Natl. Acad. Sci. USA 2012, 109, 4892–4897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Wang, X.; Schneider, A. HIF-2alpha-Mediated Activation of the Epidermal Growth Factor Receptor Potentiates Head and Neck Cancer Cell Migration in Response to Hypoxia. Carcinogenesis 2010, 31, 1202–1210. [Google Scholar] [CrossRef] [PubMed]
  208. Clarke, K.; Smith, K.; Gullick, W.J.; Harris, A.L. Mutant Epidermal Growth Factor Receptor Enhances Induction of Vascular Endothelial Growth Factor by Hypoxia and Insulin-like Growth Factor-1 via a PI3 Kinase Dependent Pathway. Br. J. Cancer 2001, 84, 1322–1329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Wouters, A.; Boeckx, C.; Vermorken, J.B.; Van den Weyngaert, D.; Peeters, M.; Lardon, F. The Intriguing Interplay between Therapies Targeting the Epidermal Growth Factor Receptor, the Hypoxic Microenvironment and Hypoxia-Inducible Factors. Curr. Pharm. Des. 2013, 19, 907–917. [Google Scholar] [CrossRef]
  210. Petit, A.M.; Rak, J.; Hung, M.C.; Rockwell, P.; Goldstein, N.; Fendly, B.; Kerbel, R.S. Neutralizing Antibodies against Epidermal Growth Factor and ErbB-2/neu Receptor Tyrosine Kinases down-Regulate Vascular Endothelial Growth Factor Production by Tumor Cells In Vitro and In Vivo: Angiogenic Implications for Signal Transduction Therapy of Solid Tumors. Am. J. Pathol. 1997, 151, 1523–1530. [Google Scholar]
  211. Riesterer, O.; Mason, K.A.; Raju, U.; Yang, Q.; Wang, L.; Hittelman, W.N.; Ang, K.K.; Milas, L. Enhanced Response to C225 of A431 Tumor Xenografts Growing in Irradiated Tumor Bed. Radiother. Oncol. 2009, 92, 383–387. [Google Scholar] [CrossRef] [PubMed]
  212. Bruns, C.J.; Harbison, M.T.; Davis, D.W.; Portera, C.A.; Tsan, R.; McConkey, D.J.; Evans, D.B.; Abbruzzese, J.L.; Hicklin, D.J.; Radinsky, R. Epidermal Growth Factor Receptor Blockade with C225 plus Gemcitabine Results in Regression of Human Pancreatic Carcinoma Growing Orthotopically in Nude Mice by Antiangiogenic Mechanisms. Clin. Cancer Res. 2000, 6, 1936–1948. [Google Scholar]
  213. Li, X.; Lu, Y.; Liang, K.; Pan, T.; Mendelsohn, J.; Fan, Z. Requirement of Hypoxia-Inducible Factor-1alpha down-Regulation in Mediating the Antitumor Activity of the Anti-Epidermal Growth Factor Receptor Monoclonal Antibody Cetuximab. Mol. Cancer Ther. 2008, 7, 1207–1217. [Google Scholar] [CrossRef] [Green Version]
  214. Li, X.; Fan, Z. The Epidermal Growth Factor Receptor Antibody Cetuximab Induces Autophagy in Cancer Cells by Downregulating HIF-1alpha and Bcl-2 and Activating the Beclin 1/hVps34 Complex. Cancer Res. 2010, 70, 5942–5952. [Google Scholar] [CrossRef] [Green Version]
  215. Lu, H.; Liang, K.; Lu, Y.; Fan, Z. The Anti-EGFR Antibody Cetuximab Sensitizes Human Head and Neck Squamous Cell Carcinoma Cells to Radiation in Part through Inhibiting Radiation-Induced Upregulation of HIF-1α. Cancer Lett. 2012, 322, 78–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Ciardiello, F.; Caputo, R.; Bianco, R.; Damiano, V.; Fontanini, G.; Cuccato, S.; De Placido, S.; Bianco, A.R.; Tortora, G. Inhibition of Growth Factor Production and Angiogenesis in Human Cancer Cells by ZD1839 (Iressa), a Selective Epidermal Growth Factor Receptor Tyrosine Kinase Inhibitor. Clin. Cancer Res. 2001, 7, 1459–1465. [Google Scholar]
  217. Hirata, A.; Ogawa, S.-I.; Kometani, T.; Kuwano, T.; Naito, S.; Kuwano, M.; Ono, M. ZD1839 (Iressa) Induces Antiangiogenic Effects through Inhibition of Epidermal Growth Factor Receptor Tyrosine Kinase. Cancer Res. 2002, 62, 2554–2560. [Google Scholar]
  218. Pore, N.; Jiang, Z.; Gupta, A.; Cerniglia, G.; Kao, G.D.; Maity, A. EGFR Tyrosine Kinase Inhibitors Decrease VEGF Expression by Both Hypoxia-Inducible Factor (HIF)-1-Independent and HIF-1-Dependent Mechanisms. Cancer Res. 2006, 66, 3197–3204. [Google Scholar] [CrossRef] [Green Version]
  219. Han, J.-Y.; Oh, S.H.; Morgillo, F.; Myers, J.N.; Kim, E.; Hong, W.K.; Lee, H.-Y. Hypoxia-Inducible Factor 1alpha and Antiangiogenic Activity of Farnesyltransferase Inhibitor SCH66336 in Human Aerodigestive Tract Cancer. J. Natl. Cancer Inst. 2005, 97, 1272–1286. [Google Scholar] [CrossRef]
  220. Catrina, S.-B.; Botusan, I.R.; Rantanen, A.; Catrina, A.I.; Pyakurel, P.; Savu, O.; Axelson, M.; Biberfeld, P.; Poellinger, L.; Brismar, K. Hypoxia-Inducible Factor-1alpha and Hypoxia-Inducible Factor-2alpha Are Expressed in Kaposi Sarcoma and Modulated by Insulin-like Growth Factor-I. Clin. Cancer Res. 2006, 12, 4506–4514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Carroll, V.A.; Ashcroft, M. Role of Hypoxia-Inducible Factor (HIF)-1alpha versus HIF-2alpha in the Regulation of HIF Target Genes in Response to Hypoxia, Insulin-like Growth Factor-I, or Loss of von Hippel-Lindau Function: Implications for Targeting the HIF Pathway. Cancer Res. 2006, 66, 6264–6270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Sutton, K.M.; Hayat, S.; Chau, N.-M.; Cook, S.; Pouyssegur, J.; Ahmed, A.; Perusinghe, N.; Le Floch, R.; Yang, J.; Ashcroft, M. Selective Inhibition of MEK1/2 Reveals a Differential Requirement for ERK1/2 Signalling in the Regulation of HIF-1 in Response to Hypoxia and IGF-1. Oncogene 2007, 26, 3920–3929. [Google Scholar] [CrossRef] [Green Version]
  223. De Francesco, E.M.; Sims, A.H.; Maggiolini, M.; Sotgia, F.; Lisanti, M.P.; Clarke, R.B. GPER Mediates the Angiocrine Actions Induced by IGF1 through the HIF-1α/VEGF Pathway in the Breast Tumor Microenvironment. Breast Cancer Res. 2017, 19, 129. [Google Scholar] [CrossRef]
  224. Tang, X.; Zhang, Q.; Shi, S.; Yen, Y.; Li, X.; Zhang, Y.; Zhou, K.; Le, A.D. Bisphosphonates Suppress Insulin-like Growth Factor 1-Induced Angiogenesis via the HIF-1alpha/VEGF Signaling Pathways in Human Breast Cancer Cells. Int. J. Cancer 2010, 126, 90–103. [Google Scholar] [CrossRef] [Green Version]
  225. Li, X.; Feng, Y.; Liu, J.; Feng, X.; Zhou, K.; Tang, X. Epigallocatechin-3-Gallate Inhibits IGF-I-Stimulated Lung Cancer Angiogenesis through Downregulation of HIF-1α and VEGF Expression. J. Nutrigenet. Nutr. 2013, 6, 169–178. [Google Scholar] [CrossRef]
  226. Tang, X.-D.; Zhou, X.; Zhou, K.-Y. Dauricine Inhibits Insulin-like Growth Factor-I-Induced Hypoxia Inducible Factor 1alpha Protein Accumulation and Vascular Endothelial Growth Factor Expression in Human Breast Cancer Cells. Acta Pharmacol. Sin. 2009, 30, 605–616. [Google Scholar] [CrossRef] [Green Version]
  227. Feldser, D.; Agani, F.; Iyer, N.V.; Pak, B.; Ferreira, G.; Semenza, G.L. Reciprocal Positive Regulation of Hypoxia-Inducible Factor 1alpha and Insulin-like Growth Factor 2. Cancer Res. 1999, 59, 3915–3918. [Google Scholar] [PubMed]
  228. Mancini, M.; Gariboldi, M.B.; Taiana, E.; Bonzi, M.C.; Craparotta, I.; Pagin, M.; Monti, E. Co-Targeting the IGF System and HIF-1 Inhibits Migration and Invasion by (triple-Negative) Breast Cancer Cells. Br. J. Cancer 2014, 110, 2865–2873. [Google Scholar] [CrossRef] [Green Version]
  229. Osher, E.; Macaulay, V.M. Therapeutic Targeting of the IGF Axis. Cells 2019, 8, 895. [Google Scholar] [CrossRef] [Green Version]
  230. Shi, Y.-H.; Wang, Y.-X.; Bingle, L.; Gong, L.-H.; Heng, W.-J.; Li, Y.; Fang, W.-G. In Vitro Study of HIF-1 Activation and VEGF Release by bFGF in the T47D Breast Cancer Cell Line under Normoxic Conditions: Involvement of PI-3K/Akt and MEK1/ERK Pathways. J. Pathol. 2005, 205, 530–536. [Google Scholar] [CrossRef]
  231. Shi, Y.-H.; Wang, Y.-X.; You, J.-F.; Heng, W.-J.; Zhong, H.-H.; Fang, W.-G. Activation of HIF-1 by bFGF in breast cancer: Role of PI-3K and MEK1/ERK pathways. Zhonghua Yi Xue Za Zhi 2004, 84, 1899–1903. [Google Scholar] [PubMed]
  232. Iqbal, S.; Zhang, S.; Driss, A.; Liu, Z.-R.; Kim, H.-R.C.; Wang, Y.; Ritenour, C.; Zhau, H.E.; Kucuk, O.; Chung, L.W.K.; et al. PDGF Upregulates Mcl-1 through Activation of β-Catenin and HIF-1α-Dependent Signaling in Human Prostate Cancer Cells. PLoS ONE 2012, 7, e30764. [Google Scholar] [CrossRef] [Green Version]
  233. Kimura, Y.; Inoue, K.; Abe, M.; Nearman, J.; Baranowska-Kortylewicz, J. PDGFRbeta and HIF-1alpha Inhibition with Imatinib and Radioimmunotherapy of Experimental Prostate Cancer. Cancer Biol. Ther. 2007, 6, 1763–1772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Schito, L.; Rey, S.; Tafani, M.; Zhang, H.; Wong, C.C.-L.; Russo, A.; Russo, M.A.; Semenza, G.L. Hypoxia-Inducible Factor 1-Dependent Expression of Platelet-Derived Growth Factor B Promotes Lymphatic Metastasis of Hypoxic Breast Cancer Cells. Proc. Natl. Acad. Sci. USA 2012, 109, E2707–E2716. [Google Scholar] [CrossRef] [Green Version]
  235. Lau, C.K.; Yang, Z.F.; Ho, D.W.; Ng, M.N.; Yeoh, G.C.T.; Poon, R.T.P.; Fan, S.T. An Akt/hypoxia-Inducible Factor-1alpha/platelet-Derived Growth Factor-BB Autocrine Loop Mediates Hypoxia-Induced Chemoresistance in Liver Cancer Cells and Tumorigenic Hepatic Progenitor Cells. Clin. Cancer Res. 2009, 15, 3462–3471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Maude, S.L.; Frey, N.; Shaw, P.A.; Aplenc, R.; Barrett, D.M.; Bunin, N.J.; Chew, A.; Gonzalez, V.E.; Zheng, Z.; Lacey, S.F.; et al. Chimeric Antigen Receptor T Cells for Sustained Remissions in Leukemia. N. Engl. J. Med. 2014, 371, 1507–1517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Turtle, C.J.; Hanafi, L.-A.; Berger, C.; Hudecek, M.; Pender, B.; Robinson, E.; Hawkins, R.; Chaney, C.; Cherian, S.; Chen, X.; et al. Immunotherapy of Non-Hodgkin’s Lymphoma with a Defined Ratio of CD8+ and CD4+ CD19-Specific Chimeric Antigen Receptor-Modified T Cells. Sci. Transl. Med. 2016, 8, 355ra116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Brentjens, R.J.; Davila, M.L.; Riviere, I.; Park, J.; Wang, X.; Cowell, L.G.; Bartido, S.; Stefanski, J.; Taylor, C.; Olszewska, M.; et al. CD19-Targeted T Cells Rapidly Induce Molecular Remissions in Adults with Chemotherapy-Refractory Acute Lymphoblastic Leukemia. Sci. Transl. Med. 2013, 5, 177ra38. [Google Scholar]
  239. Turan, T.; Kannan, D.; Patel, M.; Matthew Barnes, J.; Tanlimco, S.G.; Lu, R.; Halliwill, K.; Kongpachith, S.; Kline, D.E.; Hendrickx, W.; et al. Immune Oncology, Immune Responsiveness and the Theory of Everything. J. Immunother. Cancer 2018, 6, 50. [Google Scholar] [CrossRef] [Green Version]
  240. Bedognetti, D.; Ceccarelli, M.; Galluzzi, L.; Lu, R.; Palucka, K.; Samayoa, J.; Spranger, S.; Warren, S.; Wong, K.-K.; Ziv, E.; et al. Correction to: Toward a Comprehensive View of Cancer Immune Responsiveness: A Synopsis from the SITC Workshop. J. Immunother. Cancer 2019, 7, 167. [Google Scholar] [CrossRef] [Green Version]
  241. Pietrobon, V.; Marincola, F.M. Hypoxia and the Phenomenon of Immune Exclusion. J. Transl. Med. 2021, 19, 9. [Google Scholar] [CrossRef]
  242. Daniel, S.K.; Sullivan, K.M.; Labadie, K.P.; Pillarisetty, V.G. Hypoxia as a Barrier to Immunotherapy in Pancreatic Adenocarcinoma. Clin. Transl. Med. 2019, 8, 10. [Google Scholar] [CrossRef]
  243. Jayaprakash, P.; Ai, M.; Liu, A.; Budhani, P.; Bartkowiak, T.; Sheng, J.; Ager, C.; Nicholas, C.; Jaiswal, A.R.; Sun, Y.; et al. Targeted Hypoxia Reduction Restores T Cell Infiltration and Sensitizes Prostate Cancer to Immunotherapy. J. Clin. Investig. 2018, 128, 5137–5149. [Google Scholar] [CrossRef]
  244. Lanitis, E.; Dangaj, D.; Irving, M.; Coukos, G. Mechanisms Regulating T-Cell Infiltration and Activity in Solid Tumors. Ann. Oncol. 2017, 28, xii18–xii32. [Google Scholar] [CrossRef]
  245. Lim, A.R.; Rathmell, W.K.; Rathmell, J.C. The Tumor Microenvironment as a Metabolic Barrier to Effector T Cells and Immunotherapy. eLife 2020, 9, e55185. [Google Scholar] [CrossRef]
  246. Gillies, R.J.; Schornack, P.A.; Secomb, T.W.; Raghunand, N. Causes and Effects of Heterogeneous Perfusion in Tumors. Neoplasia 1999, 1, 197–207. [Google Scholar] [CrossRef] [Green Version]
  247. Padera, T.P.; Stoll, B.R.; Tooredman, J.B.; Capen, D.; di Tomaso, E.; Jain, R.K. Cancer Cells Compress Intratumour Vessels. Nature 2004, 427, 695. [Google Scholar] [CrossRef]
  248. Welti, J.; Loges, S.; Dimmeler, S.; Carmeliet, P. Recent Molecular Discoveries in Angiogenesis and Antiangiogenic Therapies in Cancer. J. Clin. Investig. 2013, 123, 3190–3200. [Google Scholar] [CrossRef] [Green Version]
  249. Muz, B.; de la Puente, P.; Azab, F.; Azab, A.K. The Role of Hypoxia in Cancer Progression, Angiogenesis, Metastasis, and Resistance to Therapy. Hypoxia 2015, 3, 83–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Schaaf, M.B.; Garg, A.D.; Agostinis, P. Defining the Role of the Tumor Vasculature in Antitumor Immunity and Immunotherapy. Cell Death Dis. 2018, 9, 115. [Google Scholar] [CrossRef] [Green Version]
  251. Huang, Y.; Kim, B.Y.S.; Chan, C.K.; Hahn, S.M.; Weissman, I.L.; Jiang, W. Improving Immune-Vascular Crosstalk for Cancer Immunotherapy. Nat. Rev. Immunol. 2018, 18, 195–203. [Google Scholar] [CrossRef] [PubMed]
  252. Liu, Z.; Zhao, Q.; Zheng, Z.; Liu, S.; Meng, L.; Dong, L.; Jiang, X. Vascular Normalization in Immunotherapy: A Promising Mechanisms Combined with Radiotherapy. Biomed. Pharmacother. 2021, 139, 111607. [Google Scholar] [CrossRef]
  253. Gilkes, D.M.; Bajpai, S.; Wong, C.C.; Chaturvedi, P.; Hubbi, M.E.; Wirtz, D.; Semenza, G.L. Procollagen Lysyl Hydroxylase 2 Is Essential for Hypoxia-Induced Breast Cancer Metastasis. Mol. Cancer Res. 2013, 11, 456–466. [Google Scholar] [CrossRef] [Green Version]
  254. Xiong, G.; Deng, L.; Zhu, J.; Rychahou, P.G.; Xu, R. Prolyl-4-Hydroxylase α Subunit 2 Promotes Breast Cancer Progression and Metastasis by Regulating Collagen Deposition. BMC Cancer 2014, 14, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Eisinger-Mathason, T.S.K.; Zhang, M.; Qiu, Q.; Skuli, N.; Nakazawa, M.S.; Karakasheva, T.; Mucaj, V.; Shay, J.E.S.; Stangenberg, L.; Sadri, N.; et al. Hypoxia-Dependent Modification of Collagen Networks Promotes Sarcoma Metastasis. Cancer Discov. 2013, 3, 1190–1205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Hofbauer, K.-H.; Gess, B.; Lohaus, C.; Meyer, H.E.; Katschinski, D.; Kurtz, A. Oxygen Tension Regulates the Expression of a Group of Procollagen Hydroxylases. Eur. J. Biochem. 2003, 270, 4515–4522. [Google Scholar]
  257. Aro, E.; Khatri, R.; Gerard-O’Riley, R.; Mangiavini, L.; Myllyharju, J.; Schipani, E. Hypoxia-Inducible Factor-1 (HIF-1) but Not HIF-2 Is Essential for Hypoxic Induction of Collagen Prolyl 4-Hydroxylases in Primary Newborn Mouse Epiphyseal Growth Plate Chondrocytes. J. Biol. Chem. 2012, 287, 37134–37144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Choi, J.Y.; Jang, Y.S.; Min, S.Y.; Song, J.Y. Overexpression of MMP-9 and HIF-1α in Breast Cancer Cells under Hypoxic Conditions. J. Breast Cancer 2011, 14, 88–95. [Google Scholar] [CrossRef] [Green Version]
  259. Katsuno, Y.; Lamouille, S.; Derynck, R. TGF-β Signaling and Epithelial-Mesenchymal Transition in Cancer Progression. Curr. Opin. Oncol. 2013, 25, 76–84. [Google Scholar] [CrossRef] [PubMed]
  260. Tanghetti, E.; Ria, R.; Dell’Era, P.; Urbinati, C.; Rusnati, M.; Ennas, M.G.; Presta, M. Biological Activity of Substrate-Bound Basic Fibroblast Growth Factor (FGF2): Recruitment of FGF Receptor-1 in Endothelial Cell Adhesion Contacts. Oncogene 2002, 21, 3889–3897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Erler, J.T.; Bennewith, K.L.; Cox, T.R.; Lang, G.; Bird, D.; Koong, A.; Le, Q.-T.; Giaccia, A.J. Hypoxia-Induced Lysyl Oxidase Is a Critical Mediator of Bone Marrow Cell Recruitment to Form the Premetastatic Niche. Cancer Cell 2009, 15, 35–44. [Google Scholar] [CrossRef] [Green Version]
  262. Pietras, K.; Ostman, A. Hallmarks of Cancer: Interactions with the Tumor Stroma. Exp. Cell Res. 2010, 316, 1324–1331. [Google Scholar] [PubMed]
  263. Schietke, R.; Warnecke, C.; Wacker, I.; Schödel, J.; Mole, D.R.; Campean, V.; Amann, K.; Goppelt-Struebe, M.; Behrens, J.; Eckardt, K.-U.; et al. The Lysyl Oxidases LOX and LOXL2 Are Necessary and Sufficient to Repress E-Cadherin in Hypoxia: Insights into Cellular Transformation Processes Mediated by HIF-1. J. Biol. Chem. 2010, 285, 6658–6669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Wong, C.C.-L.; Gilkes, D.M.; Zhang, H.; Chen, J.; Wei, H.; Chaturvedi, P.; Fraley, S.I.; Wong, C.-M.; Khoo, U.-S.; Ng, I.O.-L.; et al. Hypoxia-Inducible Factor 1 Is a Master Regulator of Breast Cancer Metastatic Niche Formation. Proc. Natl. Acad. Sci. USA 2011, 108, 16369–16374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Mao, Y.; Keller, E.T.; Garfield, D.H.; Shen, K.; Wang, J. Stromal Cells in Tumor Microenvironment and Breast Cancer. Cancer Metastasis Rev. 2013, 32, 303–315. [Google Scholar] [CrossRef] [Green Version]
  266. Gilkes, D.M.; Chaturvedi, P.; Bajpai, S.; Wong, C.C.; Wei, H.; Pitcairn, S.; Hubbi, M.E.; Wirtz, D.; Semenza, G.L. Collagen Prolyl Hydroxylases Are Essential for Breast Cancer Metastasis. Cancer Res. 2013, 73, 3285–3296. [Google Scholar] [CrossRef] [Green Version]
  267. Wan, R.; Mo, Y.; Chien, S.; Li, Y.; Li, Y.; Tollerud, D.J.; Zhang, Q. The Role of Hypoxia Inducible Factor-1α in the Increased MMP-2 and MMP-9 Production by Human Monocytes Exposed to Nickel Nanoparticles. Nanotoxicology 2011, 5, 568–582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Revuelta-López, E.; Castellano, J.; Roura, S.; Gálvez-Montón, C.; Nasarre, L.; Benitez, S.; Bayes-Genis, A.; Badimon, L.; Llorente-Cortés, V. Hypoxia Induces Metalloproteinase-9 Activation and Human Vascular Smooth Muscle Cell Migration Through Low-Density Lipoprotein Receptor–Related Protein 1–Mediated Pyk2 Phosphorylation. Arterioscler. Thromb. Vasc. Biol. 2013, 33, 2877–2887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  269. Kim, D.-H.; Zhou, K.; Kim, D.-K.; Park, S.; Noh, J.; Kwon, Y.; Kim, D.; Song, N.W.; Lee, J.-B.; Suh, P.-G.; et al. Analysis of Interactions between the Epidermal Growth Factor Receptor and Soluble Ligands on the Basis of Single-Molecule Diffusivity in the Membrane of Living Cells. Angew. Chem. Int. Ed. 2015, 54, 7028–7032. [Google Scholar] [CrossRef] [PubMed]
  270. Lin, C.-H.; Shih, C.-H.; Tseng, C.-C.; Yu, C.-C.; Tsai, Y.-J.; Bien, M.-Y.; Chen, B.-C. CXCL12 Induces Connective Tissue Growth Factor Expression in Human Lung Fibroblasts through the Rac1/ERK, JNK, and AP-1 Pathways. PLoS ONE 2014, 9, e104746. [Google Scholar] [CrossRef] [Green Version]
  271. Lee, Y.-L.; Cheng, W.-E.; Chen, S.-C.; Chen, C.-Y.; Shih, C.-M. The Effects of Hypoxia on the Expression of MMP-2, MMP-9 in Human Lung Adenocarcinoma A549 Cells. Eur. Respir. J. 2014, 44, P2699. [Google Scholar]
  272. Miao, J.-W.; Liu, L.-J.; Huang, J. Interleukin-6-Induced Epithelial-Mesenchymal Transition through Signal Transducer and Activator of Transcription 3 in Human Cervical Carcinoma. Int. J. Oncol. 2014, 45, 165–176. [Google Scholar] [CrossRef]
  273. Masola, V.; Carraro, A.; Granata, S.; Signorini, L.; Bellin, G.; Violi, P.; Lupo, A.; Tedeschi, U.; Onisto, M.; Gambaro, G.; et al. In Vitro Effects of Interleukin (IL)-1 Beta Inhibition on the Epithelial-to-Mesenchymal Transition (EMT) of Renal Tubular and Hepatic Stellate Cells. J. Transl. Med. 2019, 17, 12. [Google Scholar] [CrossRef] [Green Version]
  274. Long, X.; Ye, Y.; Zhang, L.; Liu, P.; Yu, W.; Wei, F.; Ren, X.; Yu, J. IL-8, a Novel Messenger to Cross-Link Inflammation and Tumor EMT via Autocrine and Paracrine Pathways (Review). Int. J. Oncol. 2016, 48, 5–12. [Google Scholar] [CrossRef] [Green Version]
  275. Palena, C.; Hamilton, D.H.; Fernando, R.I. Influence of IL-8 on the Epithelial-Mesenchymal Transition and the Tumor Microenvironment. Future Oncol. 2012, 8, 713–722. [Google Scholar] [CrossRef] [Green Version]
  276. Sistigu, A.; Di Modugno, F.; Manic, G.; Nisticò, P. Deciphering the Loop of Epithelial-Mesenchymal Transition, Inflammatory Cytokines and Cancer Immunoediting. Cytokine Growth Factor Rev. 2017, 36, 67–77. [Google Scholar] [CrossRef]
  277. Xu, J.; Lamouille, S.; Derynck, R. TGF-β-Induced Epithelial to Mesenchymal Transition. Cell Res. 2009, 19, 156–172. [Google Scholar]
  278. Gonzalez, D.M.; Medici, D. Signaling Mechanisms of the Epithelial-Mesenchymal Transition. Sci. Signal. 2014, 7, re8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Chae, Y.K.; Chang, S.; Ko, T.; Anker, J.; Agte, S.; Iams, W.; Choi, W.M.; Lee, K.; Cruz, M. Epithelial-Mesenchymal Transition (EMT) Signature Is Inversely Associated with T-Cell Infiltration in Non-Small Cell Lung Cancer (NSCLC). Sci. Rep. 2018, 8, 2918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  280. Yu, L.; Mu, Y.; Sa, N.; Wang, H.; Xu, W. Tumor Necrosis Factor α Induces Epithelial-Mesenchymal Transition and Promotes Metastasis via NF-κB Signaling Pathway-Mediated TWIST Expression in Hypopharyngeal Cancer. Oncol. Rep. 2014, 31, 321–327. [Google Scholar] [CrossRef] [Green Version]
  281. Li, C.-W.; Xia, W.; Huo, L.; Lim, S.-O.; Wu, Y.; Hsu, J.L.; Chao, C.-H.; Yamaguchi, H.; Yang, N.-K.; Ding, Q.; et al. Epithelial–Mesenchymal Transition Induced by TNF-α Requires NF-κB–Mediated Transcriptional Upregulation of Twist1. Cancer Res. 2012, 72, 1290–1300. [Google Scholar] [CrossRef] [Green Version]
  282. Hao, Y.; Baker, D.; Ten Dijke, P. TGF-β-Mediated Epithelial-Mesenchymal Transition and Cancer Metastasis. Int. J. Mol. Sci. 2019, 20, 2767. [Google Scholar] [CrossRef] [Green Version]
  283. Li, R.; Ong, S.L.; Tran, L.M.; Jing, Z.; Liu, B.; Park, S.J.; Huang, Z.L.; Walser, T.C.; Heinrich, E.L.; Lee, G.; et al. Chronic IL-1β-Induced Inflammation Regulates Epithelial-to-Mesenchymal Transition Memory Phenotypes via Epigenetic Modifications in Non-Small Cell Lung Cancer. Sci. Rep. 2020, 10, 377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Zhou, J.; Zhang, C.; Pan, J.; Chen, L.; Qi, S.-T. Interleukin-6 Induces an Epithelial-mesenchymal Transition Phenotype in Human Adamantinomatous Craniopharyngioma Cells and Promotes Tumor Cell Migration. Mol. Med. Rep. 2017, 15, 4123–4131. [Google Scholar] [CrossRef] [Green Version]
  285. Farrell, J.; Kelly, C.; Rauch, J.; Kida, K.; García-Muñoz, A.; Monsefi, N.; Turriziani, B.; Doherty, C.; Mehta, J.P.; Matallanas, D.; et al. HGF Induces Epithelial-to-Mesenchymal Transition by Modulating the Mammalian hippo/MST2 and ISG15 Pathways. J. Proteome Res. 2014, 13, 2874–2886. [Google Scholar] [CrossRef] [PubMed]
  286. Liu, F.; Song, S.; Yi, Z.; Zhang, M.; Li, J.; Yang, F.; Yin, H.; Yu, X.; Guan, C.; Liu, Y.; et al. HGF Induces EMT in Non-Small-Cell Lung Cancer through the hBVR Pathway. Eur. J. Pharmacol. 2017, 811, 180–190. [Google Scholar] [CrossRef]
  287. Strutz, F.; Zeisberg, M.; Ziyadeh, F.N.; Yang, C.-Q.; Kalluri, R.; Müller, G.A.; Neilson, E.G. Role of Basic Fibroblast Growth Factor-2 in Epithelial-Mesenchymal Transformation. Kidney Int. 2002, 61, 1714–1728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  288. Savagner, P.; Vallés, A.M.; Jouanneau, J.; Yamada, K.M.; Thiery, J.P. Alternative Splicing in Fibroblast Growth Factor Receptor 2 Is Associated with Induced Epithelial-Mesenchymal Transition in Rat Bladder Carcinoma Cells. Mol. Biol. Cell 1994, 5, 851–862. [Google Scholar] [CrossRef]
  289. Xu, Q.; Zhang, Q.; Ishida, Y.; Hajjar, S.; Tang, X.; Shi, H.; Dang, C.V.; Le, A.D. EGF Induces Epithelial-Mesenchymal Transition and Cancer Stem-like Cell Properties in Human Oral Cancer Cells via Promoting Warburg Effect. Oncotarget 2017, 8, 9557–9571. [Google Scholar] [CrossRef] [Green Version]
  290. Shu, D.; Lovicu, F.J. EGF Potentiates TGFβ-Induced Epithelial-Mesenchymal Transition (EMT) in Lens Epithelial Cells by Enhancing EGFR Signaling. Investig. Ophthalmol. Vis. Sci. 2018, 59, 1202. [Google Scholar]
  291. Wu, Q.; Hou, X.; Xia, J.; Qian, X.; Miele, L.; Sarkar, F.H.; Wang, Z. Emerging Roles of PDGF-D in EMT Progression during Tumorigenesis. Cancer Treat. Rev. 2013, 39, 640–646. [Google Scholar] [CrossRef] [Green Version]
  292. Forsythe, J.A.; Jiang, B.H.; Iyer, N.V.; Agani, F.; Leung, S.W.; Koos, R.D.; Semenza, G.L. Activation of Vascular Endothelial Growth Factor Gene Transcription by Hypoxia-Inducible Factor 1. Mol. Cell. Biol. 1996, 16, 4604–4613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Kelly, B.D.; Hackett, S.F.; Hirota, K.; Oshima, Y.; Cai, Z.; Berg-Dixon, S.; Rowan, A.; Yan, Z.; Campochiaro, P.A.; Semenza, G.L. Cell Type-Specific Regulation of Angiogenic Growth Factor Gene Expression and Induction of Angiogenesis in Nonischemic Tissue by a Constitutively Active Form of Hypoxia-Inducible Factor 1. Circ. Res. 2003, 93, 1074–1081. [Google Scholar] [CrossRef] [Green Version]
  294. Hitchon, C.; Wong, K.; Ma, G.; Reed, J.; Lyttle, D.; El-Gabalawy, H. Hypoxia-Induced Production of Stromal Cell-Derived Factor 1 (CXCL12) and Vascular Endothelial Growth Factor by Synovial Fibroblasts. Arthritis Rheum. 2002, 46, 2587–2597. [Google Scholar] [CrossRef] [PubMed]
  295. Elvert, G.; Kappel, A.; Heidenreich, R.; Englmeier, U.; Lanz, S.; Acker, T.; Rauter, M.; Plate, K.; Sieweke, M.; Breier, G.; et al. Cooperative Interaction of Hypoxia-Inducible Factor-2alpha (HIF-2alpha) and Ets-1 in the Transcriptional Activation of Vascular Endothelial Growth Factor Receptor-2 (Flk-1). J. Biol. Chem. 2003, 278, 7520–7530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Simon, M.-P.; Tournaire, R.; Pouyssegur, J. The Angiopoietin-2 Gene of Endothelial Cells Is up-Regulated in Hypoxia by a HIF Binding Site Located in Its First Intron and by the Central Factors GATA-2 and Ets-1. J. Cell. Physiol. 2008, 217, 809–818. [Google Scholar] [CrossRef]
  297. Krock, B.L.; Skuli, N.; Simon, M.C. Hypoxia-Induced Angiogenesis: Good and Evil. Genes Cancer 2011, 2, 1117–1133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  298. Chaturvedi, P.; Gilkes, D.M.; Wong, C.C.L.; Kshitiz; Luo, W.; Zhang, H.; Wei, H.; Takano, N.; Schito, L.; Levchenko, A.; et al. Hypoxia-Inducible Factor-Dependent Breast Cancer-Mesenchymal Stem Cell Bidirectional Signaling Promotes Metastasis. J. Clin. Investig. 2013, 123, 189–205. [Google Scholar] [CrossRef] [Green Version]
  299. Chen, C.; Pore, N.; Behrooz, A.; Ismail-Beigi, F.; Maity, A. Regulation of glut1 mRNA by Hypoxia-Inducible Factor-1. Interaction between H-Ras and Hypoxia. J. Biol. Chem. 2001, 276, 9519–9525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  300. Mimura, I.; Nangaku, M.; Kanki, Y.; Tsutsumi, S.; Inoue, T.; Kohro, T.; Yamamoto, S.; Fujita, T.; Shimamura, T.; Suehiro, J.-I.; et al. Dynamic Change of Chromatin Conformation in Response to Hypoxia Enhances the Expression of GLUT3 (SLC2A3) by Cooperative Interaction of Hypoxia-Inducible Factor 1 and KDM3A. Mol. Cell. Biol. 2012, 32, 3018–3032. [Google Scholar] [CrossRef] [Green Version]
  301. Wood, I.S.; Wang, B.; Lorente-Cebrián, S.; Trayhurn, P. Hypoxia Increases Expression of Selective Facilitative Glucose Transporters (GLUT) and 2-Deoxy-D-Glucose Uptake in Human Adipocytes. Biochem. Biophys. Res. Commun. 2007, 361, 468–473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Lee, H.-C.; Lin, S.-C.; Wu, M.-H.; Tsai, S.-J. Induction of Pyruvate Dehydrogenase Kinase 1 by Hypoxia Alters Cellular Metabolism and Inhibits Apoptosis in Endometriotic Stromal Cells. Reprod. Sci. 2019, 26, 734–744. [Google Scholar] [CrossRef] [PubMed]
  303. Kim, J.-W.; Tchernyshyov, I.; Semenza, G.L.; Dang, C.V. HIF-1-Mediated Expression of Pyruvate Dehydrogenase Kinase: A Metabolic Switch Required for Cellular Adaptation to Hypoxia. Cell Metab. 2006, 3, 177–185. [Google Scholar] [CrossRef] [Green Version]
  304. Semenza, G.L.; Jiang, B.-H.; Leung, S.W.; Passantino, R.; Concordet, J.-P.; 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. [Google Scholar] [CrossRef] [Green Version]
  305. Semenza, G.L.; Roth, P.H.; Fang, H.M.; Wang, G.L. Transcriptional Regulation of Genes Encoding Glycolytic Enzymes by Hypoxia-Inducible Factor 1. J. Biol. Chem. 1994, 269, 23757–23763. [Google Scholar] [CrossRef]
  306. Miranda-Gonçalves, V.; Granja, S.; Martinho, O.; Honavar, M.; Pojo, M.; Costa, B.M.; Pires, M.M.; Pinheiro, C.; Cordeiro, M.; Bebiano, G.; et al. Hypoxia-Mediated Upregulation of MCT1 Expression Supports the Glycolytic Phenotype of Glioblastomas. Oncotarget 2016, 7, 46335–46353. [Google Scholar] [CrossRef] [Green Version]
  307. Jamali, S.; Klier, M.; Ames, S.; Barros, L.F.; McKenna, R.; Deitmer, J.W.; Becker, H.M. Hypoxia-Induced Carbonic Anhydrase IX Facilitates Lactate Flux in Human Breast Cancer Cells by Non-Catalytic Function. Sci. Rep. 2015, 5, 13605. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  308. Saraswati, S.; Guo, Y.; Atkinson, J.; Young, P.P. Prolonged Hypoxia Induces Monocarboxylate Transporter-4 Expression in Mesenchymal Stem Cells Resulting in a Secretome That Is Deleterious to Cardiovascular Repair. Stem Cells 2015, 33, 1333–1344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  309. Brahimi-Horn, M.C.; Pouysségur, J. Hypoxia in Cancer Cell Metabolism and pH Regulation. Essays Biochem. 2007, 43, 165–178. [Google Scholar] [PubMed] [Green Version]
  310. Chiche, J.; Brahimi-Horn, M.C.; Pouysségur, J. Tumour Hypoxia Induces a Metabolic Shift Causing Acidosis: A Common Feature in Cancer. J. Cell. Mol. Med. 2010, 14, 771–794. [Google Scholar] [CrossRef] [Green Version]
  311. Svastová, E.; Hulíková, A.; Rafajová, M.; Zat’ovicová, M.; Gibadulinová, A.; Casini, A.; Cecchi, A.; Scozzafava, A.; Supuran, C.T.; Pastorek, J.; et al. Hypoxia Activates the Capacity of Tumor-Associated Carbonic Anhydrase IX to Acidify Extracellular pH. FEBS Lett. 2004, 577, 439–445. [Google Scholar] [CrossRef] [Green Version]
  312. Bobulescu, I.A.; Di Sole, F.; Moe, O.W. Na+/H+ Exchangers: Physiology and Link to Hypertension and Organ Ischemia. Curr. Opin. Nephrol. Hypertens. 2005, 14, 485–494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  313. Sedlakova, O.; Svastova, E.; Takacova, M.; Kopacek, J.; Pastorek, J.; Pastorekova, S. Carbonic Anhydrase IX, a Hypoxia-Induced Catalytic Component of the pH Regulating Machinery in Tumors. Front. Physiol. 2014, 4, 400. [Google Scholar] [CrossRef] [Green Version]
  314. Hussain, S.A.; Ganesan, R.; Reynolds, G.; Gross, L.; Stevens, A.; Pastorek, J.; Murray, P.G.; Perunovic, B.; Anwar, M.S.; Billingham, L.; et al. Hypoxia-Regulated Carbonic Anhydrase IX Expression Is Associated with Poor Survival in Patients with Invasive Breast Cancer. Br. J. Cancer 2007, 96, 104–109. [Google Scholar] [CrossRef]
  315. Parks, S.K.; Pouyssegur, J. The Na+/HCO3 Co-Transporter SLC4A4 Plays a Role in Growth and Migration of Colon and Breast Cancer Cells. J. Cell. Physiol. 2015, 230, 1954–1963. [Google Scholar] [CrossRef]
  316. Ye, Z.; Yue, L.; Shi, J.; Shao, M.; Wu, T. Role of IDO and TDO in Cancers and Related Diseases and the Therapeutic Implications. J. Cancer 2019, 10, 2771–2782. [Google Scholar] [CrossRef] [Green Version]
  317. Xiang, L.; Mou, J.; Shao, B.; Wei, Y.; Liang, H.; Takano, N.; Semenza, G.L.; Xie, G. Glutaminase 1 Expression in Colorectal Cancer Cells Is Induced by Hypoxia and Required for Tumor Growth, Invasion, and Metastatic Colonization. Cell Death Dis. 2019, 10, 40. [Google Scholar] [CrossRef] [Green Version]
  318. Chimote, A.A.; Kuras, Z.; Conforti, L. Disruption of kv1.3 Channel Forward Vesicular Trafficking by Hypoxia in Human T Lymphocytes. J. Biol. Chem. 2012, 287, 2055–2067. [Google Scholar] [CrossRef] [Green Version]
  319. Conforti, L.; Petrovic, M.; Mohammad, D.; Lee, S.; Ma, Q.; Barone, S.; Filipovich, A.H. Hypoxia Regulates Expression and Activity of Kv1.3 Channels in T Lymphocytes: A Possible Role in T Cell Proliferation. J. Immunol. 2003, 170, 695–702. [Google Scholar] [CrossRef] [Green Version]
  320. Rodriguez, P.C.; Ernstoff, M.S.; Hernandez, C.; Atkins, M.; Zabaleta, J.; Sierra, R.; Ochoa, A.C. Arginase I-Producing Myeloid-Derived Suppressor Cells in Renal Cell Carcinoma Are a Subpopulation of Activated Granulocytes. Cancer Res. 2009, 69, 1553–1560. [Google Scholar] [CrossRef] [Green Version]
  321. Rodríguez, P.C.; Ochoa, A.C. Arginine Regulation by Myeloid Derived Suppressor Cells and Tolerance in Cancer: Mechanisms and Therapeutic Perspectives. Immunol. Rev. 2008, 222, 180–191. [Google Scholar] [CrossRef] [PubMed]
  322. Platten, M.; Wick, W.; Van den Eynde, B.J. Tryptophan Catabolism in Cancer: Beyond IDO and Tryptophan Depletion. Cancer Res. 2012, 72, 5435–5440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Ziani, L.; Chouaib, S.; Thiery, J. Alteration of the Antitumor Immune Response by Cancer-Associated Fibroblasts. Front. Immunol. 2018, 9, 414. [Google Scholar] [CrossRef]
  324. Schmidt, S.K.; Ebel, S.; Keil, E.; Woite, C.; Ernst, J.F.; Benzin, A.E.; Rupp, J.; Däubener, W. Regulation of IDO Activity by Oxygen Supply: Inhibitory Effects on Antimicrobial and Immunoregulatory Functions. PLoS ONE 2013, 8, e63301. [Google Scholar] [CrossRef] [Green Version]
  325. Mohapatra, S.R.; Sadik, A.; Tykocinski, L.-O.; Dietze, J.; Poschet, G.; Heiland, I.; Opitz, C.A. Hypoxia Inducible Factor 1α Inhibits the Expression of Immunosuppressive Tryptophan-2,3-Dioxygenase in Glioblastoma. Front. Immunol. 2019, 10, 2762. [Google Scholar] [CrossRef]
  326. Sener, Z.; Cederkvist, F.H.; Volchenkov, R.; Holen, H.L.; Skålhegg, B.S. T Helper Cell Activation and Expansion Is Sensitive to Glutaminase Inhibition under Both Hypoxic and Normoxic Conditions. PLoS ONE 2016, 11, e0160291. [Google Scholar] [CrossRef] [Green Version]
  327. Du, R.; Lu, K.V.; Petritsch, C.; Liu, P.; Ganss, R.; Passegué, E.; Song, H.; Vandenberg, S.; Johnson, R.S.; Werb, Z.; et al. HIF1alpha Induces the Recruitment of Bone Marrow-Derived Vascular Modulatory Cells to Regulate Tumor Angiogenesis and Invasion. Cancer Cell 2008, 13, 206–220. [Google Scholar] [CrossRef] [Green Version]
  328. Lin, S.; Wan, S.; Sun, L.; Hu, J.; Fang, D.; Zhao, R.; Yuan, S.; Zhang, L. Chemokine C-C Motif Receptor 5 and C-C Motif Ligand 5 Promote Cancer Cell Migration under Hypoxia. Cancer Sci. 2012, 103, 904–912. [Google Scholar] [CrossRef]
  329. Schioppa, T.; Uranchimeg, B.; Saccani, A.; Biswas, S.K.; Doni, A.; Rapisarda, A.; Bernasconi, S.; Saccani, S.; Nebuloni, M.; Vago, L.; et al. Regulation of the Chemokine Receptor CXCR4 by Hypoxia. J. Exp. Med. 2003, 198, 1391–1402. [Google Scholar] [CrossRef] [Green Version]
  330. Casazza, A.; Laoui, D.; Wenes, M.; Rizzolio, S.; Bassani, N.; Mambretti, M.; Deschoemaeker, S.; Van Ginderachter, J.A.; Tamagnone, L.; Mazzone, M. Impeding Macrophage Entry into Hypoxic Tumor Areas by Sema3A/Nrp1 Signaling Blockade Inhibits Angiogenesis and Restores Antitumor Immunity. Cancer Cell 2013, 24, 695–709. [Google Scholar] [CrossRef] [Green Version]
  331. Dineen, S.P.; Lynn, K.D.; Holloway, S.E.; Miller, A.F.; Sullivan, J.P.; Shames, D.S.; Beck, A.W.; Barnett, C.C.; Fleming, J.B.; Brekken, R.A. Vascular Endothelial Growth Factor Receptor 2 Mediates Macrophage Infiltration into Orthotopic Pancreatic Tumors in Mice. Cancer Res. 2008, 68, 4340–4346. [Google Scholar] [CrossRef] [Green Version]
  332. Grimshaw, M.J.; Naylor, S.; Balkwill, F.R. Endothelin-2 Is a Hypoxia-Induced Autocrine Survival Factor for Breast Tumor Cells. Mol. Cancer Ther. 2002, 1, 1273–1281. [Google Scholar]
  333. Grimshaw, M.J. Endothelins and Hypoxia-Inducible Factor in Cancer. Endocr. Relat. Cancer 2007, 14, 233–244. [Google Scholar] [CrossRef]
  334. Leek, R.D.; Hunt, N.C.; Landers, R.J.; Lewis, C.E.; Royds, J.A.; Harris, A.L. Macrophage Infiltration Is Associated with VEGF and EGFR Expression in Breast Cancer. J. Pathol. 2000, 190, 430–436. [Google Scholar] [CrossRef]
  335. Poth, J.M.; Brodsky, K.; Ehrentraut, H.; Grenz, A.; Eltzschig, H.K. Transcriptional Control of Adenosine Signaling by Hypoxia-Inducible Transcription Factors during Ischemic or Inflammatory Disease. J. Mol. Med. 2013, 91, 183–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Synnestvedt, K.; Furuta, G.T.; Comerford, K.M.; Louis, N.; Karhausen, J.; Eltzschig, H.K.; Hansen, K.R.; Thompson, L.F.; Colgan, S.P. Ecto-5′-Nucleotidase (CD73) Regulation by Hypoxia-Inducible Factor-1 Mediates Permeability Changes in Intestinal Epithelia. J. Clin. Investig. 2002, 110, 993–1002. [Google Scholar] [CrossRef] [PubMed]
  337. Zhang, H.; Lu, H.; Xiang, L.; Bullen, J.W.; Zhang, C.; Samanta, D.; Gilkes, D.M.; He, J.; Semenza, G.L. HIF-1 Regulates CD47 Expression in Breast Cancer Cells to Promote Evasion of Phagocytosis and Maintenance of Cancer Stem Cells. Proc. Natl. Acad. Sci. USA 2015, 112, E6215–E6223. [Google Scholar] [CrossRef] [Green Version]
  338. Huang, C.-Y.; Ye, Z.-H.; Huang, M.-Y.; Lu, J.-J. Regulation of CD47 Expression in Cancer Cells. Transl. Oncol. 2020, 13, 100862. [Google Scholar] [CrossRef]
  339. Jiang, Y.; Zhu, Y.; Wang, X.; Gong, J.; Hu, C.; Guo, B.; Zhu, B.; Li, Y. Temporal Regulation of HIF-1 and NF-κB in Hypoxic Hepatocarcinoma Cells. Oncotarget 2015, 6, 9409–9419. [Google Scholar] [CrossRef] [Green Version]
  340. Chandel, N.S.; Trzyna, W.C.; McClintock, D.S.; Schumacker, P.T. Role of Oxidants in NF-Kappa B Activation and TNF-Alpha Gene Transcription Induced by Hypoxia and Endotoxin. J. Immunol. 2000, 165, 1013–1021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  341. Rius, J.; Guma, M.; Schachtrup, C.; Akassoglou, K.; Zinkernagel, A.S.; Nizet, V.; Johnson, R.S.; Haddad, G.G.; Karin, M. NF-kappaB Links Innate Immunity to the Hypoxic Response through Transcriptional Regulation of HIF-1alpha. Nature 2008, 453, 807–811. [Google Scholar] [CrossRef] [Green Version]
  342. Belaiba, R.S.; Bonello, S.; Zähringer, C.; Schmidt, S.; Hess, J.; Kietzmann, T.; Görlach, 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  343. Pawlus, M.R.; Wang, L.; Hu, C.-J. STAT3 and HIF1α Cooperatively Activate HIF1 Target Genes in MDA-MB-231 and RCC4 Cells. Oncogene 2014, 33, 1670–1679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  344. Jung, J.E.; Lee, H.G.; Cho, I.H.; Chung, D.H.; Yoon, S.-H.; Yang, Y.M.; Lee, J.W.; Choi, S.; Park, J.-W.; Ye, S.-K.; et al. STAT3 Is a Potential Modulator of HIF-1-Mediated VEGF Expression in Human Renal Carcinoma Cells. FASEB J. 2005, 19, 1296–1298. [Google Scholar] [CrossRef]
  345. Yaguchi, T.; Goto, Y.; Kido, K.; Mochimaru, H.; Sakurai, T.; Tsukamoto, N.; Kudo-Saito, C.; Fujita, T.; Sumimoto, H.; Kawakami, Y. Immune Suppression and Resistance Mediated by Constitutive Activation of Wnt/β-Catenin Signaling in Human Melanoma Cells. J. Immunol. 2012, 189, 2110–2117. [Google Scholar] [CrossRef] [Green Version]
  346. Spranger, S.; Bao, R.; Gajewski, T.F. Melanoma-Intrinsic β-Catenin Signalling Prevents Anti-Tumour Immunity. Nature 2015, 523, 231–235. [Google Scholar] [CrossRef]
  347. Zhang, Z.; Yao, L.; Yang, J.; Wang, Z.; Du, G. PI3K/Akt and HIF-1 Signaling Pathway in Hypoxia-ischemia (Review). Mol. Med. Rep. 2018, 18, 3547–3554. [Google Scholar] [CrossRef] [Green Version]
  348. Sethumadhavan, S.; Silva, M.; Philbrook, P.; Nguyen, T.; Hatfield, S.M.; Ohta, A.; Sitkovsky, M.V. Hypoxia and Hypoxia-Inducible Factor (HIF) Downregulate Antigen-Presenting MHC Class I Molecules Limiting Tumor Cell Recognition by T Cells. PLoS ONE 2017, 12, e0187314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  349. Rankin, E.B.; Fuh, K.C.; Castellini, L.; Viswanathan, K.; Finger, E.C.; Diep, A.N.; LaGory, E.L.; Kariolis, M.S.; Chan, A.; Lindgren, D.; et al. Direct Regulation of GAS6/AXL Signaling by HIF Promotes Renal Metastasis through SRC and MET. Proc. Natl. Acad. Sci. USA 2014, 111, 13373–13378. [Google Scholar] [CrossRef] [Green Version]
  350. Zhu, D.; Wang, Y.; Singh, I.; Bell, R.D.; Deane, R.; Zhong, Z.; Sagare, A.; Winkler, E.A.; Zlokovic, B.V. Protein S Controls Hypoxic/ischemic Blood-Brain Barrier Disruption through the TAM Receptor Tyro3 and Sphingosine 1-Phosphate Receptor. Blood 2010, 115, 4963–4972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  351. Zhu, Y.-Z.; Wang, W.; Xian, N.; Wu, B. Inhibition of TYRO3/Akt Signaling Participates in Hypoxic Injury in Hippocampal Neurons. Neural Regen. Res. 2016, 11, 752. [Google Scholar] [CrossRef] [PubMed]
  352. Mishra, A.; Wang, J.; Shiozawa, Y.; McGee, S.; Kim, J.; Jung, Y.; Joseph, J.; Berry, J.E.; Havens, A.; Pienta, K.J.; et al. Hypoxia Stabilizes GAS6/Axl Signaling in Metastatic Prostate Cancer. Mol. Cancer Res. 2012, 10, 703–712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  353. Noman, M.Z.; Desantis, G.; Janji, B.; Hasmim, M.; Karray, S.; Dessen, P.; Bronte, V.; Chouaib, S. PD-L1 Is a Novel Direct Target of HIF-1α, and Its Blockade under Hypoxia Enhanced MDSC-Mediated T Cell Activation. J. Exp. Med. 2014, 211, 781–790. [Google Scholar] [CrossRef] [PubMed]
  354. Cubillos-Zapata, C.; Avendaño-Ortiz, J.; Hernandez-Jimenez, E.; Toledano, V.; Casas-Martin, J.; Varela-Serrano, A.; Torres, M.; Almendros, I.; Casitas, R.; Fernández-Navarro, I.; et al. Hypoxia-Induced PD-L1/PD-1 Crosstalk Impairs T-Cell Function in Sleep Apnoea. Eur. Respir. J. 2017, 50, 1700833. [Google Scholar] [CrossRef]
  355. Wen, Q.; Han, T.; Wang, Z.; Jiang, S. Role and Mechanism of Programmed Death-Ligand 1 in Hypoxia-Induced Liver Cancer Immune Escape. Oncol. Lett. 2020, 19, 2595–2601. [Google Scholar] [CrossRef] [Green Version]
  356. Lequeux, A.; Noman, M.Z.; Xiao, M.; Sauvage, D.; Van Moer, K.; Viry, E.; Bocci, I.; Hasmim, M.; Bosseler, M.; Berchem, G.; et al. Impact of Hypoxic Tumor Microenvironment and Tumor Cell Plasticity on the Expression of Immune Checkpoints. Cancer Lett. 2019, 458, 13–20. [Google Scholar] [CrossRef]
  357. Corpechot, C.; Barbu, V.; Wendum, D.; Kinnman, N.; Rey, C.; Poupon, R.; Housset, C.; Rosmorduc, O. Hypoxia-Induced VEGF and Collagen I Expressions Are Associated with Angiogenesis and Fibrogenesis in Experimental Cirrhosis. Hepatology 2002, 35, 1010–1021. [Google Scholar] [CrossRef]
  358. Gilkes, D.M.; Semenza, G.L.; Wirtz, D. Hypoxia and the Extracellular Matrix: Drivers of Tumour Metastasis. Nat. Rev. Cancer 2014, 14, 430–439. [Google Scholar] [CrossRef] [Green Version]
  359. Gilkes, D.M.; Bajpai, S.; Chaturvedi, P.; Wirtz, D.; Semenza, G.L. Hypoxia-Inducible Factor 1 (HIF-1) Promotes Extracellular Matrix Remodeling under Hypoxic Conditions by Inducing P4HA1, P4HA2, and PLOD2 Expression in Fibroblasts. J. Biol. Chem. 2013, 288, 10819–10829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  360. McKay, T.B.; Hjortdal, J.; Priyadarsini, S.; Karamichos, D. Acute Hypoxia Influences Collagen and Matrix Metalloproteinase Expression by Human Keratoconus Cells In Vitro. PLoS ONE 2017, 12, e0176017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  361. Bauer, A.T.; Bürgers, H.F.; Rabie, T.; Marti, H.H. Matrix Metalloproteinase-9 Mediates Hypoxia-Induced Vascular Leakage in the Brain via Tight Junction Rearrangement. J. Cereb. Blood Flow Metab. 2010, 30, 837–848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  362. Zhu, S.; Zhou, Y.; Wang, L.; Zhang, J.; Wu, H.; Xiong, J.; Zhang, J.; Tian, Y.; Wang, C.; Wu, H. Transcriptional Upregulation of MT2-MMP in Response to Hypoxia Is Promoted by HIF-1α in Cancer Cells. Mol. Carcinog. 2011, 50, 770–780. [Google Scholar] [CrossRef]
  363. Kuczek, D.E.; Larsen, A.M.H.; Thorseth, M.-L.; Carretta, M.; Kalvisa, A.; Siersbæk, M.S.; Simões, A.M.C.; Roslind, A.; Engelholm, L.H.; Noessner, E.; et al. Collagen Density Regulates the Activity of Tumor-Infiltrating T Cells. J Immunother Cancer 2019, 7, 68. [Google Scholar] [CrossRef] [Green Version]
  364. Casey, T.M.; Eneman, J.; Crocker, A.; White, J.; Tessitore, J.; Stanley, M.; Harlow, S.; Bunn, J.Y.; Weaver, D.; Muss, H.; et al. Cancer Associated Fibroblasts Stimulated by Transforming Growth Factor beta1 (TGF-Beta 1) Increase Invasion Rate of Tumor Cells: A Population Study. Breast Cancer Res. Treat. 2008, 110, 39–49. [Google Scholar] [CrossRef] [PubMed]
  365. Yoon, H.; Tang, C.-M.; Banerjee, S.; Delgado, A.L.; Yebra, M.; Davis, J.; Sicklick, J.K. TGF-β1-Mediated Transition of Resident Fibroblasts to Cancer-Associated Fibroblasts Promotes Cancer Metastasis in Gastrointestinal Stromal Tumor. Oncogenesis 2021, 10, 13. [Google Scholar] [CrossRef]
  366. McMahon, S.; Charbonneau, M.; Grandmont, S.; Richard, D.E.; Dubois, C.M. Transforming Growth Factor beta1 Induces Hypoxia-Inducible Factor-1 Stabilization through Selective Inhibition of PHD2 Expression. J. Biol. Chem. 2006, 281, 24171–24181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  367. Mingyuan, X.; Qianqian, P.; Shengquan, X.; Chenyi, Y.; Rui, L.; Yichen, S.; Jinghong, X. Hypoxia-Inducible Factor-1α Activates Transforming Growth Factor-β1/Smad Signaling and Increases Collagen Deposition in Dermal Fibroblasts. Oncotarget 2018, 9, 3188–3197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  368. Brabletz, T.; Kalluri, R.; Nieto, M.A.; Weinberg, R.A. EMT in Cancer. Nat. Rev. Cancer 2018, 18, 128–134. [Google Scholar] [CrossRef]
  369. Yang, P.; Hu, Y.; Zhou, Q. The CXCL12-CXCR4 Signaling Axis Plays a Key Role in Cancer Metastasis and Is a Potential Target for Developing Novel Therapeutics against Metastatic Cancer. Curr. Med. Chem. 2020, 27, 5543–5561. [Google Scholar] [CrossRef]
  370. Sullivan, N.J.; Sasser, A.K.; Axel, A.E.; Vesuna, F.; Raman, V.; Ramirez, N.; Oberyszyn, T.M.; Hall, B.M. Interleukin-6 Induces an Epithelial–mesenchymal Transition Phenotype in Human Breast Cancer Cells. Oncogene 2009, 28, 2940–2947. [Google Scholar] [CrossRef] [Green Version]
  371. Borthwick, L.A. The IL-1 Cytokine Family and Its Role in Inflammation and Fibrosis in the Lung. Semin. Immunopathol. 2016, 38, 517–534. [Google Scholar] [CrossRef] [Green Version]
  372. Suarez-Carmona, M.; Lesage, J.; Cataldo, D.; Gilles, C. EMT and Inflammation: Inseparable Actors of Cancer Progression. Mol. Oncol. 2017, 11, 805–823. [Google Scholar] [CrossRef]
  373. Said, N.A.B.M.; Williams, E.D. Growth Factors in Induction of Epithelial-Mesenchymal Transition and Metastasis. Cells Tissues Organs 2011, 193, 85–97. [Google Scholar] [CrossRef] [PubMed]
  374. Wang, L.; Saci, A.; Szabo, P.M.; Chasalow, S.D.; Castillo-Martin, M.; Domingo-Domenech, J.; Siefker-Radtke, A.; Sharma, P.; Sfakianos, J.P.; Gong, Y.; et al. EMT- and Stroma-Related Gene Expression and Resistance to PD-1 Blockade in Urothelial Cancer. Nat. Commun. 2018, 9, 3503. [Google Scholar] [CrossRef]
  375. Hugo, W.; Zaretsky, J.M.; Sun, L.; Song, C.; Moreno, B.H.; Hu-Lieskovan, S.; Berent-Maoz, B.; Pang, J.; Chmielowski, B.; Cherry, G.; et al. Genomic and Transcriptomic Features of Response to Anti-PD-1 Therapy in Metastatic Melanoma. Cell 2016, 165, 35–44. [Google Scholar] [CrossRef] [Green Version]
  376. Du, Y.; Miao, W.; Jiang, X.; Cao, J.; Wang, B.; Wang, Y.; Yu, J.; Wang, X.; Liu, H. The Epithelial to Mesenchymal Transition Related Gene Calumenin Is an Adverse Prognostic Factor of Bladder Cancer Correlated with Tumor Microenvironment Remodeling, Gene Mutation, and Ferroptosis. Front. Oncol. 2021, 11, 683951. [Google Scholar] [CrossRef]
  377. Lou, Y.; Diao, L.; Cuentas, E.R.P.; Denning, W.L.; Chen, L.; Fan, Y.H.; Byers, L.A.; Wang, J.; Papadimitrakopoulou, V.A.; Behrens, C.; et al. Epithelial-Mesenchymal Transition Is Associated with a Distinct Tumor Microenvironment Including Elevation of Inflammatory Signals and Multiple Immune Checkpoints in Lung Adenocarcinoma. Clin. Cancer Res. 2016, 22, 3630–3642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  378. Seiler, R.; Ashab, H.A.D.; Erho, N.; van Rhijn, B.W.G.; Winters, B.; Douglas, J.; Van Kessel, K.E.; Fransen van de Putte, E.E.; Sommerlad, M.; Wang, N.Q.; et al. Impact of Molecular Subtypes in Muscle-Invasive Bladder Cancer on Predicting Response and Survival after Neoadjuvant Chemotherapy. Eur. Urol. 2017, 72, 544–554. [Google Scholar] [CrossRef] [PubMed]
  379. DeBerardinis, R.J.; Chandel, N.S. We Need to Talk about the Warburg Effect. Nat. Metab. 2020, 2, 127–129. [Google Scholar] [CrossRef] [PubMed]
  380. Varum, S.; Rodrigues, A.S.; Moura, M.B.; Momcilovic, O.; Easley, C.A., 4th; Ramalho-Santos, J.; Van Houten, B.; Schatten, G. Energy Metabolism in Human Pluripotent Stem Cells and Their Differentiated Counterparts. PLoS ONE 2011, 6, e20914. [Google Scholar] [CrossRef] [Green Version]
  381. Xie, J.; Wu, H.; Dai, C.; Pan, Q.; Ding, Z.; Hu, D.; Ji, B.; Luo, Y.; Hu, X. Beyond Warburg Effect--Dual Metabolic Nature of Cancer Cells. Sci. Rep. 2014, 4, 4927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  382. Wang, J.; Ye, C.; Chen, C.; Xiong, H.; Xie, B.; Zhou, J.; Chen, Y.; Zheng, S.; Wang, L. Glucose Transporter GLUT1 Expression and Clinical Outcome in Solid Tumors: A Systematic Review and Meta-Analysis. Oncotarget 2017, 8, 16875–16886. [Google Scholar] [CrossRef] [Green Version]
  383. Kierans, S.J.; Taylor, C.T. Regulation of Glycolysis by the Hypoxia-Inducible Factor (HIF): Implications for Cellular Physiology. J. Physiol. 2021, 599, 23–37. [Google Scholar] [CrossRef]
  384. Lum, J.J.; Bui, T.; Gruber, M.; Gordan, J.D.; DeBerardinis, R.J.; Covello, K.L.; Simon, M.C.; Thompson, C.B. The Transcription Factor HIF-1alpha Plays a Critical Role in the Growth Factor-Dependent Regulation of Both Aerobic and Anaerobic Glycolysis. Genes Dev. 2007, 21, 1037–1049. [Google Scholar] [CrossRef] [Green Version]
  385. Erra Díaz, F.; Dantas, E.; Geffner, J. Unravelling the Interplay between Extracellular Acidosis and Immune Cells. Mediat. Inflamm. 2018, 2018, 1218297. [Google Scholar] [CrossRef]
  386. Kellum, J.A.; Song, M.; Li, J. Science Review: Extracellular Acidosis and the Immune Response: Clinical and Physiologic Implications. Crit. Care 2004, 8, 331–336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  387. Reshkin, S.J.; Cardone, R.A.; Harguindey, S. Na+-H+ Exchanger, pH Regulation and Cancer. Recent Pat. Anticancer Drug Discov. 2013, 8, 85–99. [Google Scholar] [CrossRef]
  388. Geiger, R.; Rieckmann, J.C.; Wolf, T.; Basso, C.; Feng, Y.; Fuhrer, T.; Kogadeeva, M.; Picotti, P.; Meissner, F.; Mann, M.; et al. L-Arginine Modulates T Cell Metabolism and Enhances Survival and Anti-Tumor Activity. Cell 2016, 167, 829–842.e13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  389. Rodriguez, P.C.; Quiceno, D.G.; Ochoa, A.C. L-Arginine Availability Regulates T-Lymphocyte Cell-Cycle Progression. Blood 2007, 109, 1568–1573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  390. Thomas, L.W.; Ashcroft, M. Exploring the Molecular Interface between Hypoxia-Inducible Factor Signalling and Mitochondria. Cell. Mol. Life Sci. 2019, 76, 1759–1777. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  391. Nakaya, M.; Xiao, Y.; Zhou, X.; Chang, J.-H.; Chang, M.; Cheng, X.; Blonska, M.; Lin, X.; Sun, S.-C. Inflammatory T Cell Responses Rely on Amino Acid Transporter ASCT2 Facilitation of Glutamine Uptake and mTORC1 Kinase Activation. Immunity 2014, 40, 692–705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  392. Sinclair, L.V.; Rolf, J.; Emslie, E.; Shi, Y.-B.; Taylor, P.M.; Cantrell, D.A. Control of Amino-Acid Transport by Antigen Receptors Coordinates the Metabolic Reprogramming Essential for T Cell Differentiation. Nat. Immunol. 2013, 14, 500–508. [Google Scholar] [CrossRef] [Green Version]
  393. Vito, A.; El-Sayes, N.; Mossman, K. Hypoxia-Driven Immune Escape in the Tumor Microenvironment. Cells 2020, 9, 992. [Google Scholar] [CrossRef]
  394. Chiu, D.K.-C.; Tse, A.P.-W.; Xu, I.M.-J.; Di Cui, J.; Lai, R.K.-H.; Li, L.L.; Koh, H.-Y.; Tsang, F.H.-C.; Wei, L.L.; Wong, C.-M.; et al. Hypoxia Inducible Factor HIF-1 Promotes Myeloid-Derived Suppressor Cells Accumulation through ENTPD2/CD39L1 in Hepatocellular Carcinoma. Nat. Commun. 2017, 8, 517. [Google Scholar] [CrossRef] [Green Version]
  395. Abou Khouzam, R.; Brodaczewska, K.; Filipiak, A.; Zeinelabdin, N.A.; Buart, S.; Szczylik, C.; Kieda, C.; Chouaib, S. Tumor Hypoxia Regulates Immune Escape/Invasion: Influence on Angiogenesis and Potential Impact of Hypoxic Biomarkers on Cancer Therapies. Front. Immunol. 2020, 11, 613114. [Google Scholar] [CrossRef] [PubMed]
  396. Lemke, G. Biology of the TAM Receptors. Cold Spring Harb. Perspect. Biol. 2013, 5, a009076. [Google Scholar] [CrossRef] [PubMed]
  397. Michaels, A.D.; Newhook, T.E.; Adair, S.J.; Morioka, S.; Goudreau, B.J.; Nagdas, S.; Mullen, M.G.; Persily, J.B.; Bullock, T.N.J.; Slingluff, C.L., Jr.; et al. CD47 Blockade as an Adjuvant Immunotherapy for Resectable Pancreatic Cancer. Clin. Cancer Res. 2018, 24, 1415–1425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  398. Soto-Pantoja, D.R.; Terabe, M.; Ghosh, A.; Ridnour, L.A.; DeGraff, W.G.; Wink, D.A.; Berzofsky, J.A.; Roberts, D.D. CD47 in the Tumor Microenvironment Limits Cooperation between Antitumor T-Cell Immunity and Radiotherapy. Cancer Res. 2014, 74, 6771–6783. [Google Scholar] [CrossRef] [Green Version]
  399. Kaur, S.; Chang, T.; Singh, S.P.; Lim, L.; Mannan, P.; Garfield, S.H.; Pendrak, M.L.; Soto-Pantoja, D.R.; Rosenberg, A.Z.; Jin, S.; et al. CD47 Signaling Regulates the Immunosuppressive Activity of VEGF in T Cells. J. Immunol. 2014, 193, 3914–3924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  400. Samanta, D.; Park, Y.; Ni, X.; Li, H.; Zahnow, C.A.; Gabrielson, E.; Pan, F.; Semenza, G.L. Chemotherapy Induces Enrichment of CD47+/CD73+/PDL1+ Immune Evasive Triple-Negative Breast Cancer Cells. Proc. Natl. Acad. Sci. USA 2018, 115, E1239–E1248. [Google Scholar] [CrossRef] [Green Version]
  401. Li, Y.; Patel, S.P.; Roszik, J.; Qin, Y. Hypoxia-Driven Immunosuppressive Metabolites in the Tumor Microenvironment: New Approaches for Combinational Immunotherapy. Front. Immunol. 2018, 9, 1591. [Google Scholar] [CrossRef] [Green Version]
  402. Chambers, A.M.; Matosevic, S. Immunometabolic Dysfunction of Natural Killer Cells Mediated by the Hypoxia-CD73 Axis in Solid Tumors. Front Mol. Biosci. 2019, 6, 60. [Google Scholar] [CrossRef] [Green Version]
  403. Ahmad, A.; Ahmad, S.; Glover, L.; Miller, S.M.; Shannon, J.M.; Guo, X.; Franklin, W.A.; Bridges, J.P.; Schaack, J.B.; Colgan, S.P.; et al. Adenosine A2A Receptor Is a Unique Angiogenic Target of HIF-2α in Pulmonary Endothelial Cells. Proc. Natl. Acad. Sci. USA 2009, 106, 10684–10689. [Google Scholar] [CrossRef] [Green Version]
  404. Lan, J.; Lu, H.; Samanta, D.; Salman, S.; Lu, Y.; Semenza, G.L. Hypoxia-Inducible Factor 1-Dependent Expression of Adenosine Receptor 2B Promotes Breast Cancer Stem Cell Enrichment. Proc. Natl. Acad. Sci. USA 2018, 115, E9640–E9648. [Google Scholar] [CrossRef] [Green Version]
  405. Steingold, J.M.; Hatfield, S.M. Targeting Hypoxia-A2A Adenosinergic Immunosuppression of Antitumor T Cells During Cancer Immunotherapy. Front. Immunol. 2020, 11, 570041. [Google Scholar] [CrossRef]
  406. Hatfield, S.M.; Kjaergaard, J.; Lukashev, D.; Belikoff, B.; Schreiber, T.H.; Sethumadhavan, S.; Abbott, R.; Philbrook, P.; Thayer, M.; Shujia, D.; et al. Systemic Oxygenation Weakens the Hypoxia and Hypoxia Inducible Factor 1α-Dependent and Extracellular Adenosine-Mediated Tumor Protection. J. Mol. Med. 2014, 92, 1283–1292. [Google Scholar] [CrossRef] [PubMed]
  407. Hatfield, S.M.; Sitkovsky, M. A2A Adenosine Receptor Antagonists to Weaken the Hypoxia-HIF-1α Driven Immunosuppression and Improve Immunotherapies of Cancer. Curr. Opin. Pharmacol. 2016, 29, 90–96. [Google Scholar] [CrossRef] [Green Version]
  408. Murthy, A.; Gerber, S.A.; Koch, C.J.; Lord, E.M. Intratumoral Hypoxia Reduces IFN-γ-Mediated Immunity and MHC Class I Induction in a Preclinical Tumor Model. Immunohorizons 2019, 3, 149–160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  409. Vaupel, P.; Mayer, A. Hypoxia in Cancer: Significance and Impact on Clinical Outcome. Cancer Metastasis Rev. 2007, 26, 225–239. [Google Scholar] [CrossRef] [PubMed]
  410. Walsh, J.C.; Lebedev, A.; Aten, E.; Madsen, K.; Marciano, L.; Kolb, H.C. The Clinical Importance of Assessing Tumor Hypoxia: Relationship of Tumor Hypoxia to Prognosis and Therapeutic Opportunities. Antioxid. Redox Signal. 2014, 21, 1516–1554. [Google Scholar] [CrossRef]
  411. Foehrenbacher, A.; Secomb, T.W.; Wilson, W.R.; Hicks, K.O. Design of Optimized Hypoxia-Activated Prodrugs Using Pharmacokinetic/pharmacodynamic Modeling. Front. Oncol. 2013, 3, 314. [Google Scholar] [CrossRef] [Green Version]
  412. Phillips, R.M. Targeting the Hypoxic Fraction of Tumours Using Hypoxia-Activated Prodrugs. Cancer Chemother. Pharmacol. 2016, 77, 441–457. [Google Scholar] [CrossRef] [Green Version]
  413. Hamis, S.; Kohandel, M.; Dubois, L.J.; Yaromina, A.; Lambin, P.; Powathil, G.G. Combining Hypoxia-Activated Prodrugs and Radiotherapy in Silico: Impact of Treatment Scheduling and the Intra-Tumoural Oxygen Landscape. PLoS Comput. Biol. 2020, 16, e1008041. [Google Scholar] [CrossRef]
  414. Spiegelberg, L.; Houben, R.; Niemans, R.; de Ruysscher, D.; Yaromina, A.; Theys, J.; Guise, C.P.; Smaill, J.B.; Patterson, A.V.; Lambin, P.; et al. Hypoxia-Activated Prodrugs and (lack of) Clinical Progress: The Need for Hypoxia-Based Biomarker Patient Selection in Phase III Clinical Trials. Clin. Transl. Radiat. Oncol. 2019, 15, 62–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  415. Walker, P.R. Let There Be Oxygen and T Cells. J. Clin. Investig. 2018, 128, 4761–4763. [Google Scholar] [CrossRef] [PubMed]
  416. Damgaci, S.; Enriquez-Navas, P.M.; Pilon-Thomas, S.; Guvenis, A.; Gillies, R.J.; Ibrahim-Hashim, A. Immunotherapy on Acid: Opportunities and Challenges. Eur. J. Clin. Nutr. 2020, 74, 3–6. [Google Scholar] [CrossRef]
  417. Ding, X.; Yu, J.; Hu, M. The Relationship between Expression of PD-L1 and HIF-1α in Glioma Cells under Hypoxia. J. Clin. Orthod. 2021, 39, e14043. [Google Scholar]
  418. Choi, W.S.W.; Boland, J.; Lin, J. Hypoxia-Inducible Factor-2α as a Novel Target in Renal Cell Carcinoma. J Kidney Cancer VHL 2021, 8, 1–7. [Google Scholar] [CrossRef]
  419. Tatli Dogan, H.; Kiran, M.; Bilgin, B.; Kiliçarslan, A.; Sendur, M.A.N.; Yalçin, B.; Ardiçoglu, A.; Atmaca, A.F.; Gumuskaya, B. Prognostic Significance of the Programmed Death Ligand 1 Expression in Clear Cell Renal Cell Carcinoma and Correlation with the Tumor Microenvironment and Hypoxia-Inducible Factor Expression. Diagn. Pathol. 2018, 13, 60. [Google Scholar] [CrossRef]
  420. Magnussen, A.L.; Mills, I.G. Vascular Normalisation as the Stepping Stone into Tumour Microenvironment Transformation. Br. J. Cancer 2021, 125, 324–336. [Google Scholar] [CrossRef]
  421. Goel, S.; Wong, A.H.-K.; Jain, R.K. Vascular Normalization as a Therapeutic Strategy for Malignant and Nonmalignant Disease. Cold Spring Harb. Perspect. Med. 2012, 2, a006486. [Google Scholar] [CrossRef]
  422. Huang, Y.; Goel, S.; Duda, D.G.; Fukumura, D.; Jain, R.K. Vascular Normalization as an Emerging Strategy to Enhance Cancer Immunotherapy. Cancer Res. 2013, 73, 2943–2948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  423. Huang, Y.; Yuan, J.; Righi, E.; Kamoun, W.S.; Ancukiewicz, M.; Nezivar, J.; Santosuosso, M.; Martin, J.D.; Martin, M.R.; Vianello, F.; et al. Vascular Normalizing Doses of Antiangiogenic Treatment Reprogram the Immunosuppressive Tumor Microenvironment and Enhance Immunotherapy. Proc. Natl. Acad. Sci. USA 2012, 109, 17561–17566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  424. Dewhirst, M.W.; Mowery, Y.M.; Mitchell, J.B.; Cherukuri, M.K.; Secomb, T.W. Rationale for Hypoxia Assessment and Amelioration for Precision Therapy and Immunotherapy Studies. J. Clin. Investig. 2019, 129, 489–491. [Google Scholar] [CrossRef]
  425. He, H.; Liao, Q.; Zhao, C.; Zhu, C.; Feng, M.; Liu, Z.; Jiang, L.; Zhang, L.; Ding, X.; Yuan, M.; et al. Conditioned CAR-T Cells by Hypoxia-Inducible Transcription Amplification (HiTA) System Significantly Enhances Systemic Safety and Retains Antitumor Efficacy. J. Immunother. Cancer 2021, 9, e002755. [Google Scholar] [CrossRef] [PubMed]
  426. Kosti, P.; Opzoomer, J.W.; Larios-Martinez, K.I.; Henley-Smith, R.; Scudamore, C.L.; Okesola, M.; Taher, M.Y.M.; Davies, D.M.; Muliaditan, T.; Larcombe-Young, D.; et al. Hypoxia-Sensing CAR T Cells Provide Safety and Efficacy in Treating Solid Tumors. Cell. Rep. Med. 2021, 2, 100227. [Google Scholar] [CrossRef]
  427. Gropper, Y.; Feferman, T.; Shalit, T.; Salame, T.-M.; Porat, Z.; Shakhar, G. Culturing CTLs under Hypoxic Conditions Enhances Their Cytolysis and Improves Their Anti-Tumor Function. Cell Rep. 2017, 20, 2547–2555. [Google Scholar] [CrossRef] [Green Version]
  428. Caldwell, C.C.; Kojima, H.; Lukashev, D.; Armstrong, J.; Farber, M.; Apasov, S.G.; Sitkovsky, M.V. Differential Effects of Physiologically Relevant Hypoxic Conditions on T Lymphocyte Development and Effector Functions. J. Immunol. 2001, 167, 6140–6149. [Google Scholar] [CrossRef]
  429. Vuillefroy de Silly, R.; Ducimetière, L.; Yacoub Maroun, C.; Dietrich, P.-Y.; Derouazi, M.; Walker, P.R. Phenotypic Switch of CD8+ T Cells Reactivated under Hypoxia toward IL-10 Secreting, Poorly Proliferative Effector Cells. Eur. J. Immunol. 2015, 45, 2263–2275. [Google Scholar] [CrossRef] [Green Version]
  430. Roman, J.; Rangasamy, T.; Guo, J.; Sugunan, S.; Meednu, N.; Packirisamy, G.; Shimoda, L.A.; Golding, A.; Semenza, G.; Georas, S.N. T-Cell Activation under Hypoxic Conditions Enhances IFN-Gamma Secretion. Am. J. Respir. Cell Mol. Biol. 2010, 42, 123–128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  431. Veliça, P.; Cunha, P.P.; Vojnovic, N.; Foskolou, I.P.; Bargiela, D.; Gojkovic, M.; Rundqvist, H.; Johnson, R.S. Modified Hypoxia-Inducible Factor Expression in CD8+ T Cells Increases Antitumor Efficacy. Cancer Immunol. Res. 2021, 9, 401–414. [Google Scholar] [CrossRef] [PubMed]
  432. Berahovich, R.; Liu, X.; Zhou, H.; Tsadik, E.; Xu, S.; Golubovskaya, V.; Wu, L. Hypoxia Selectively Impairs CAR-T Cells In Vitro. Cancers 2019, 11, 602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  433. Vuillefroy de Silly, R.; Dietrich, P.-Y.; Walker, P.R. Hypoxia and Antitumor CD8+ T Cells: An Incompatible Alliance? Oncoimmunology 2016, 5, e1232236. [Google Scholar] [CrossRef] [Green Version]
  434. Nakagawa, Y.; Negishi, Y.; Shimizu, M.; Takahashi, M.; Ichikawa, M.; Takahashi, H. Effects of Extracellular pH and Hypoxia on the Function and Development of Antigen-Specific Cytotoxic T Lymphocytes. Immunol. Lett. 2015, 167, 72–86. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic of tumor growth and progression (affected by hormone and GF signaling): somatic mutation, clonal expansion, intraluminal cell proliferation and intraluminal lesion formation, invasion, dissemination, the formation of micrometastases, resistant tumor clones, angiogenesis, and ultimately metastases.
Figure 1. Schematic of tumor growth and progression (affected by hormone and GF signaling): somatic mutation, clonal expansion, intraluminal cell proliferation and intraluminal lesion formation, invasion, dissemination, the formation of micrometastases, resistant tumor clones, angiogenesis, and ultimately metastases.
Cancers 14 00539 g001
Figure 2. In physioxia (right), hypoxia-inducible factor (HIF) prolyl hydroxylase domain enzymes (PHDs) regulate the stability of HIF proteins by post-translational hydroxylation of two conserved prolyl residues in its α-subunit in an oxygen-dependent manner. Hydroxylation of HIF creates a binding site for pVHL that directs the polyubiquitylation of HIF-1α and its proteasomal degradation. In hypoxic conditions (left), HIF-1α binds to HIF-1β to form a heterodimer that acts as transcription factor, upregulating a variety of genes.
Figure 2. In physioxia (right), hypoxia-inducible factor (HIF) prolyl hydroxylase domain enzymes (PHDs) regulate the stability of HIF proteins by post-translational hydroxylation of two conserved prolyl residues in its α-subunit in an oxygen-dependent manner. Hydroxylation of HIF creates a binding site for pVHL that directs the polyubiquitylation of HIF-1α and its proteasomal degradation. In hypoxic conditions (left), HIF-1α binds to HIF-1β to form a heterodimer that acts as transcription factor, upregulating a variety of genes.
Cancers 14 00539 g002
Figure 3. (A) Schematic of HypoxiCAR T-cell activation: A hypoxia-inducible CAR construct, whose transcription is mediated by HIF-1. (↑ means increase, ↓ means decrease) (B) HypoxiCAR T-cells are not excluded from HIF-1-stabilized regions of the tumor. Immunofluorescence images from a human oral tongue carcinoma. Nuclei are stained with DAPI (blue), anti-CD3 antibody (green), and anti-HIF-1α antibody (red); white denotes CD3 and HIF-1α co-localization. Scale bar 100 μm. (Reprinted with permission from Elsevier. Copyright (2021) Cell Reports Medicine [426]).
Figure 3. (A) Schematic of HypoxiCAR T-cell activation: A hypoxia-inducible CAR construct, whose transcription is mediated by HIF-1. (↑ means increase, ↓ means decrease) (B) HypoxiCAR T-cells are not excluded from HIF-1-stabilized regions of the tumor. Immunofluorescence images from a human oral tongue carcinoma. Nuclei are stained with DAPI (blue), anti-CD3 antibody (green), and anti-HIF-1α antibody (red); white denotes CD3 and HIF-1α co-localization. Scale bar 100 μm. (Reprinted with permission from Elsevier. Copyright (2021) Cell Reports Medicine [426]).
Cancers 14 00539 g003
Table 1. Hormone and GF-dependent regulation of HIF expression/signaling (↑is increase, ↓is decrease).
Table 1. Hormone and GF-dependent regulation of HIF expression/signaling (↑is increase, ↓is decrease).
SignalingCancerResponseReference
ERα-mediated estrogen signalingOvarian↑ HIF-1α protein/signaling[180]
Thyroid↑ HIF-1α protein[183]
Breast↑ HIF-1α protein/signaling[184]
Breast↓ HIF-2α mRNA/protein[186]
Breast↑ HIF-1α mRNA[192]
GPER-mediated estrogen signalingBreast, CAFs↑ HIF-1α mRNA/protein/signaling[155]
EGF/EGFR signalingProstate↑ HIF-1α protein/signaling[199]
NSCLC↑ HIF-1α signaling[200]
Breast↑ HIF-1α protein/signaling[201,203]
Colorectal↑ HIF-1α mRNA/protein[202]
HeregulinBreast↑ HIF-1α protein/signaling[174]
IGF-IColon↑ HIF-1α protein/signaling[175]
NSCLC, HNSCC↑ HIF-1α protein/signaling[219]
Kaposi sarcoma↑ HIF-1α and HIF-2α protein/signaling[220]
Breast↑ HIF-1α and HIF-2α protein/signaling[221,222,223]
bFGFBreast↑ HIF-1α protein/signaling[176,230,231]
Table 2. Determinants of immune exclusion that are influenced by hypoxia.
Table 2. Determinants of immune exclusion that are influenced by hypoxia.
Physical BarriersImpediments to Direct Contact Between T-Cells and Cancer CellsReference
Stromal fibrosisEpidermal growth factor (EGF), platelet-derived growth factor (PDGF), fibroblast growth factor 2 (FGF2), CXCL12, TGF-β, zinc finger E-box binding homeobox 1 and 2 (ZEB1, ZEB2) proteins, Snail, Slug, Twist, Goosecoid, FOXC2, LOX, PLOD1, PLOD2, P4HA1, P4HA2, MMP2, and MMP9[253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271]
Epithelial-mesenchymal transition (EMT)Tumor necrosis factor α (TNF-α), TGF-β, interleukin 1 (IL-1), interleukin-6 (IL-6) and interleukin-8 (IL-8), hepatocyte growth factor (HGF), basic fibroblast growth factor (bFGF), epidermal growth factor (EGF), platelet-derived growth factor (PDGF)[272,273,274,275,276,277,278,279,280,281,282,283,284,285,286,287,288,289,290,291]
Vascular accessVEGF-family, angiopoietin-2 (Ang-2), transforming growth factor beta (TGF-β), platelet-derived growth factor B (PDGFB), placental growth factor (PGF), connective tissue growth factor (CTGF), stem cell factor (SCF), stromal cell-derived factor 1 (CXCL12), leptin, endoglin, nitric oxide synthase 2, haemoxygenase-1, endothelin-1 (ET-1), VEGF receptor-2, endothelial receptor tyrosine kinase (Tie-2)[83,234,292,293,294,295,296,297,298]
Functional BarriersBiological or metabolic interactions between cancer, stromal and immune cells limiting migration, function, and/or survival of T-cells
Metabolic barriersWarburg effect: enzymes glucose transporters (GLUTs 1–3), pyruvate dehydrogenase kinase 1 (PDK1), lactic dehydrogenase A (LDHA) and pyruvate kinase M2 subtype (PKM2), mono-carboxylate transporters (MCTs)[299,300,301,302,303,304,305,306,307,308]
Acidification of the TME: carbonic anhydrases (CAs), Na+/H+ exchanger (NHE1), bicarbonate transporters (SLC4A4)[90,309,310,311,312,313,314,315]
Amino acids depletion and ionic misbalance: indoleamine 2,3 dioxygenase (IDO), Glutaminase 1 (GLS1), Kv1.3[316,317,318,319,320,321,322,323,324,325,326]
Soluble factors and “Don’t eat-me” signalsMyeloid cells recruitment and activity: CCL5, CXCL12, CXCR4, VEGF, Sema3A, CCL28, endothelin 1 and 2, TGF-β[327,328,329,330,331,332,333,334]
Adenosine signaling: CD39, CD73[335,336]
“Don’t eat me” signals: CD47/signal regulatory protein (SIRP)-α axis[337,338]
Tumor cell-intrinsic signalingPathways involved in immune escape: extended PI3K pathway signaling, β-catenin/signaling, STAT-3 activation, MAPK signaling, p53 signaling[199,339,340,341,342,343,344,345,346,347]
Downregulation of molecules necessary for effector immune cells recognition: major histocompatibility class-I (MHC-I)[348]
TAM receptor tyrosine kinases: Tyro3, Axl and Mertk[349,350,351,352]
Dynamic barriersInteractions between cancer and T-cells resulting in limited function
Checkpoint/ligand interactions: upregulation of CTLA-4, PD-L1, HLA-G[353,354,355,356]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lappano, R.; Todd, L.A.; Stanic, M.; Cai, Q.; Maggiolini, M.; Marincola, F.; Pietrobon, V. Multifaceted Interplay between Hormones, Growth Factors and Hypoxia in the Tumor Microenvironment. Cancers 2022, 14, 539. https://doi.org/10.3390/cancers14030539

AMA Style

Lappano R, Todd LA, Stanic M, Cai Q, Maggiolini M, Marincola F, Pietrobon V. Multifaceted Interplay between Hormones, Growth Factors and Hypoxia in the Tumor Microenvironment. Cancers. 2022; 14(3):539. https://doi.org/10.3390/cancers14030539

Chicago/Turabian Style

Lappano, Rosamaria, Lauren A. Todd, Mia Stanic, Qi Cai, Marcello Maggiolini, Francesco Marincola, and Violena Pietrobon. 2022. "Multifaceted Interplay between Hormones, Growth Factors and Hypoxia in the Tumor Microenvironment" Cancers 14, no. 3: 539. https://doi.org/10.3390/cancers14030539

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop