Next Article in Journal
Directional Persistence of Cell Migration in Schizophrenia Patient-Derived Olfactory Cells
Next Article in Special Issue
Endothelium-Derived Hyperpolarizing Factor (EDHF) Mediates Acetylsalicylic Acid (Aspirin) Vasodilation of Pregnant Rat Mesenteric Arteries
Previous Article in Journal
Changes in Sphingolipid Profile of Benzo[a]pyrene-Transformed Human Bronchial Epithelial Cells Are Reflected in the Altered Composition of Sphingolipids in Their Exosomes
Previous Article in Special Issue
Bisphenol a Interferes with Uterine Artery Features and Impairs Rat Feto-Placental Growth
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hypoxia Pathway Proteins and Their Impact on the Blood Vasculature

Institute of Clinical Chemistry and Laboratory Medicine, Technische Universität Dresden, 01307 Dresden, Germany
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(17), 9191; https://doi.org/10.3390/ijms22179191
Submission received: 28 July 2021 / Revised: 16 August 2021 / Accepted: 21 August 2021 / Published: 25 August 2021
(This article belongs to the Special Issue Molecular Vascular Physiology)

Abstract

:
Every cell in the body requires oxygen for its functioning, in virtually every animal, and a tightly regulated system that balances oxygen supply and demand is therefore fundamental. The vascular network is one of the first systems to sense oxygen, and deprived oxygen (hypoxia) conditions automatically lead to a cascade of cellular signals that serve to circumvent the negative effects of hypoxia, such as angiogenesis associated with inflammation, tumor development, or vascular disorders. This vascular signaling is driven by central transcription factors, namely the hypoxia inducible factors (HIFs), which determine the expression of a growing number of genes in endothelial cells and pericytes. HIF functions are tightly regulated by oxygen sensors known as the HIF-prolyl hydroxylase domain proteins (PHDs), which are enzymes that hydroxylate HIFs for eventual proteasomal degradation. HIFs, as well as PHDs, represent attractive therapeutic targets under various pathological settings, including those involving vascular (dys)function. We focus on the characteristics and mechanisms by which vascular cells respond to hypoxia under a variety of conditions.

1. Introduction

Maintaining oxygen (O2) homeostasis is crucial for the survival of many species and the principal transcription factors that regulate the cellular response to low O2 tension, or hypoxia, are the hypoxia-inducible factors (HIFs). Multiple HIF isoforms exist but HIF1α and HIF2α are the most studied, and both bind to HIFβ to form the functional heterodimer. Whilst HIFβ is expressed constitutively, HIFα protein expression is regulated by HIF-prolyl hydroxylase domain (PHD) enzymes and the factor inhibiting HIF (FIH). Specifically, PHDs hydroxylate HIFα subunits at two specific proline residues in an oxygen- and iron-dependent manner; this promotes their ubiquitination by the VHL (von-Hippel–Lindau) enzyme and ultimate degradation by the 26S proteasome. As PHDs require oxygen for catalytic action, they are rendered inactive under hypoxic conditions, which allows HIFα stabilization and its translocation into the nucleus, where it interacts with CBP/p300 and HIFβ, and binds to hypoxia responsive elements (HREs) in target genes that typically promote greater transcription. Of note, even though HIF1 and HIF2 share several target genes, they also have unique binding profiles. For example, whereas HIF1 is mainly linked to regulation of metabolic programming, HIF2 modulates angiogenic extracellular signaling, guidance cues, and extracellular matrix (ECM) remodeling factors [1,2]. Further, their protein levels peak at different times during hypoxia, e.g., while the HIF1 maximum appears at around 12 h followed by a gradual decrease, HIF2 shows delayed induction followed by a stable plateau; thus, HIF2 might be better associated with chronic hypoxia. In contrast to PHDs, FIH hydroxylates HIFα at an asparagine residue, which also prevents its interaction with HIFβ and other factors (e.g., CBP/p300) and inhibits subsequent transcriptional activity [3,4,5]. To date, three main PHD isoforms have been identified, namely, PHDs 1–3; they are expressed in distinct tissues, have specific subcellular localization patterns, and preferentially bind to specific HIFα subunits. Further, deletion of each PHD results in characteristic developmental phenotypes, and as only systemic PHD2 deletion is embryonic lethal, it is arguably the most prominent oxygen sensor. Interestingly, the criticality of oxygen sensing is underscored by the fact that PHD2 transcription itself is also induced by HIF1, which not only ensures swift removal of HIF after sufficient oxygenation has been restored [3,6] but also creates a feedback loop that allows tight regulation of genes participating in the hypoxia response. PHDs also display HIF-independent effects, e.g., whilst PHD1 can hydroxylate IKKβ and thus modulate the activity of nuclear factor kappa B (NFκB), a major inflammatory regulator, PHD2 can hydroxylate eukaryotic elongation factor 2 kinase (eEF2K), which controls protein synthesis [6]. Furthermore, even though combined genomic data analysis using mathematical modelling has identified 6000 hypoxia responsive genes, about 70% of these had no HRE in their proximal promoter [7], suggesting that most of the effects of hypoxia are either HIF-independent or are indirectly targeted by HIFs (see Figure 1).
Endothelial cells (ECs) are a very central cell type in virtually every organ of the body as they are in direct contact with the blood stream; thus, it is intuitive that they are also very responsive to hypoxia, and hence HIFs, in which capacity they regulate vital processes, including cell survival, growth, cell invasion, and energy metabolism. Specifically, while ECs exposed to chronic hypoxia pivot towards increased glycolysis, biosynthesis of amino acids, carbon metabolism, pentose phosphate pathway, fructose/mannose, and cysteine/methionine metabolism [23], those subjected to acute hypoxia exhibit upregulation of genes involved in pyruvate metabolism and glucose transport, suggesting higher occurrence of glycolysis in hypoxic ECs [24]. Besides these direct effects of hypoxia on ECs, many indirect results, such as lactate accumulation in the tumor environment, have also been documented, and accruing evidence suggests that glycolytic activity in ECs, along with metabolism in general, might drive angiogenesis [25,26,27]. The purpose of this review is to discuss our current understanding of the role of hypoxia pathway proteins (HPPs) during angiogenesis in both physiological as well as pathological conditions, with a particular focus on cancer. Additionally, we address present therapeutic strategies targeting HPPs, bringing special attention to the intricate regulatory mechanisms of the HPPs and their context-dependent role.

2. Cardiovascular System Formation in the Embryo

Vasculogenesis is the de novo creation of blood vessels by migration and differentiation of endothelial progenitor cells, or angioblasts; in contrast, angiogenesis refers to the formation of new blood vessels that branch out from pre-existing vessels through sprouting and intussusceptive microvascular growth [28]. During embryogenesis, vasculogenesis generates new blood vessels but their remodeling occurs through angiogenesis [29,30]; consequently, angiogenesis is more prevalent in adults than vasculogenesis.
During embryonic vasculogenesis, hemangioblasts migrate to the yolk sac, where they eventually associate with each other to form blood islands and differentiate into endothelial cells, hematopoietic cells, and vascular smooth muscle cells (SMCs). Hemangioblasts on the outer regions of the blood islands are called angioblasts or precursors of endothelial cells. Fibroblast growth factor 2 (FGF-2) is a major regulator of both vasculogenesis and angiogenesis, and studies in avian embryo have implicated a role for FGF-2 during angioblast differentiation [31,32,33]. Interestingly, HIF1α regulates FGF2 protein levels [34,35] and vice versa, creating a feedback loop [34]. In mouse embryos, FGF-2 induces the expression of vascular endothelial growth factor (VEGF) receptor-2 (VEGFR-2, also known as Flk1), the earliest known receptor for VEGF and a well-characterized activator of angiogenesis [36]. VEGF family members act as chemoattractants for angioblasts, regulate cell proliferation, tube formation, and vessel branching, and are endothelial cell-survival factors; VEGF-A influences angioblast differentiation and initiates a series of steps that lead to a mature vascular network [37]. Hypoxia plays a role in the regulation of this pathway, as it can differentially induce or repress the expression of VEGF family members and their receptors, depending on cell type [38,39,40].
Rapid growth of the embryo during development consumes large amounts of oxygen, leading to hypoxia and increased HIF expression. Thus, in mice while HIF1α expression is high at E8.5 and increases between E9.5 to E18, HIFβ expression varies by less than twofold during E8.5 to E18 [41]. Further, as HIF1α and HIF2α require HIFβ to form their respective functional complexes, deletion of HIFβ dampens the transcriptional activity of both proteins and produces a wide array of phenotypes, including hematopoietic abnormalities and death due to vascular defects [42]. Additionally, embryoid body (EB) cells show lower VEGF and erythropoietin (EPO) expression—these factors are targets of HIF1α that play a crucial role in vasculogenesis and angiogenesis. Additionally, these HIFβ deletion mutants have fewer hematopoietic stem and progenitor cells (HSPCs) at E8.5 and E9.5 and display greater apoptosis in hematopoietic cells. Interestingly, this last phenotype could be rescued by the administration of VEGF [43]. Vessels in the HIFβ−/− embryos were also disorganized, especially in the para-aortic splanchnopleura (pSp)/aorta-gonad-mesonephros region, and HIFβ deletion in pSp cultures inhibited both vasculogenesis and angiogenesis. This phenotype was rescued by adding Sca-1+ hematopoietic cells or VEGF to these cultures, suggesting that HIFs coordinate early EC emergence and vessel development by regulating hematopoietic cell survival and production of growth factors by these cells [43].
All solid organs contain monocyte-derived cells that appear early in organogenesis, and for many organs, these cells are critical for normal development. For example, mouse embryonic macrophages were found to have pro-proliferative gene expression signatures, irrespective of their tissue of origin, linking them to tumor-associated macrophages [44]. Macrophages can be found between days 8 and 9.5 post-coitum in both extra-embryonic and embryonic tissues, where they associate with angiogenic tip cells; notably, ablation of these macrophages resulted in fewer vessel intersections in the hindbrain [45]. Moreover, HIF1 deficient zebrafish display impaired macrophage mobilization from the aorta-gonad-mesonephros (AGM) region [46].
As with HIFβ deletion mutants, heterozygous loss of HIF2α is embryonic lethal in most mice due to cardiac, vascular, and neural malformations, and mice that grow to adulthood exhibit hypocellularity in the bone marrow, fewer red and white blood cells [47], reduced myeloid multi-lineage progenitors, committed erythroid progenitors, and hemoglobin content in erythroid colonies [48,49,50]. Likewise, specific deletion of HIF1α in embryonic cells expressing VE-cadherin led to a deficiency of HSPCs in the aorta and the placenta [51], but the effects of deleting either HIF1α or HIF2α in the placenta were less severe than homozygous deletions or absence of HIFβ [52]. Furthermore, hypoxic cells expressing high levels of HIF-1α were present in the placenta, while HIF2α expression was mainly found in the decidua. Notably, no significant tissue hypoxia was detected in the placental labyrinth where eNOS was upregulated, but inhibition of NOS resulted in ubiquitous placental hypoxia [53]. Thus, hypoxia appears to affect placental vascularization through tight regulation of HIF1α, HIF2α, and HIFβ.

3. Structure of a Blood Vessel

There are three major types of blood vessels, viz., arteries that carry blood away from the heart, the branching and merging capillaries, and veins that return deoxygenated blood back to the heart. The walls of both arteries and veins consist of three layers: (1) the tunica intima, which is the innermost layer that is in contact with the lumen, and hence, blood flow; (2) the tunica media, an intermediate layer; and (3) the tunica adventitia (or Externa), the outermost layer. The tunica intima is attached to a basal lamina and consists of a single layer of ECs that is in direct contact with the blood stream. The tunica media mostly consists of elastin and smooth muscle cells (SMCs), while the tunica externa consists of a collagen-rich ECM produced by myofibroblasts. Mature blood vessels contain resident progenitor cells capable of differentiating into SMCs that repopulate the tunica media and the tunica intima. They can also produce a unique ECM that varies during vessel maturation, which is responsible for the elasticity and other mechanical properties of these vessels. Notably, hypoxia regulates the expression of many growth factors and pro-enzymes present in the ECM, some of which are also implicated in the degradation of the ECM. Specifically, HIF1α can induce the expression of matrix metalloproteases (MMP) -2, -9, and -15, whereas HIF2α can induce MMP14, and hypoxic environments can downregulate tissue inhibitors of metalloproteases (TIMP) -2 and -3, which generally function as inhibitors of MMPs. Thus, HIFs modulate many characteristics of the ECM, including composition, posttranslational modifications, and rearrangements [54,55,56].
As one of the major functions of blood vessels is to transport oxygen that is bound to red blood cells to different parts of the body, insufficient oxygen pressure induces vasodilation to promote blood flow and thereby enhance tissue oxygenation. This corrective mechanism is partially dependent on hypoxia-induced upregulation of endothelial nitric oxide synthase (eNOS), and the consequent increase in NO production relaxes the smooth muscle cells, leading to vasodilation and increased blood flow [27,57]. ECs can interact with the surrounding ECM through integrins, which have multiple ligands. Sites where integrins are clustered can act as focal adhesion sites between the ECM and various cell types, including platelets, endothelial cells, leukocytes, and smooth muscle cells. Although integrins mainly bind to components in the ECM and to cell-surface ligands, they can also bind to some soluble ligands, such as chemokines or cytokines [58,59]. As the expression of numerous integrins is controlled by the hypoxia pathway, conceivably, variations in oxygen levels can modulate cell–cell and ECM–cell interactions. For example, HIF can regulate integrin alpha IIb beta 3 expression in platelets, leading to altered platelet aggregation under hypoxic conditions [60]. Furthermore, variations in oxygen can also alter the production of red blood cells. During homeostasis in adults, erythropoiesis takes place in the bone marrow. This process is controlled by erythropoietin, which is strictly regulated by changes in oxygen partial pressure and HIFs. A more in-depth review of this process was recently published by our research group [61].

4. Angiogenesis

Angiogenesis typically starts in response to the presence of angiogenic cytokines, most noticeably VEGF, and many factors can induce the release of these cytokines, including wounding, ischemia, and hypoxia [62]. Angiogenesis is a complex process and many of the associated factors can play multiple distinct roles, depending on the context in which they are released. Moreover, while cytokines are primary effectors, multiple receptor families also regulate guidance during vascular morphogenesis and are often driven by hypoxia (see Table 1), thereby underscoring the importance of hypoxia as a factor that regulates vascular patterning [63,64]. Therefore, several major proteins involved in angiogenesis initiation are regulated by HIF1, HIF2, or both wherein, while deletion of HIF1α results in inhibition of angiogenesis, absence of HIF2α results in increased angiogenic sprouting but reduced perfusion [23,65,66]. Some of the key ligand–receptor couples involved in the regulation of angiogenesis are VEGFA/VEGFR, Dll4/Notch, ANG2/Tie2, Wnt/Frizzled, PDGF/PDGFR, and ECM/integrins [67].
An excellent and well-described example of hypoxia-driven angiogenesis pertains to the retina of mice in the first weeks after birth. In the initial step of sprouting angiogenesis, hypoxic astrocytes in the regions of the retina that show no vascularization express VEGF, which then binds to VEGFR2 on endothelial tip cells [55]. Concurrently, ANG2/TIE2 interaction eases the vessel wall [115], reducing endothelial cell–vascular smooth muscle cell interactions (EC-SMC). The ECs then become activated in response to angiogenic signals that are released in response to hypoxia, including VEGF, ANG2, and FGF (see Figure 2). Hypoxia also induces the expression of metalloproteinases that can cause degradation of the existing ECM, which in turn undergoes structural modifications to morph into a provisional ECM; importantly, these changes lead to the release of several growth factors and proenzymes embedded in the ECM [54,55].
Next, endothelial stalk cells proliferate and elongate the new branch while trailing behind the tip cell; they also establish tight and adherent junctions, ensuring structural integrity of the new sprouts (Figure 2). Tip cells lead the new branch from the front and extend the long and dynamic actin-based filopodia in response to VEGF, sensing the hypoxic microenvironment and the VEGF gradient through their filopodia and migrating towards the hypoxic areas [118,119,120]. VEGF and hypoxia promote the expression of Dll4 in endothelial cell membranes, which regulates tip/stalk specification by binding to Notch1 expressed in stalk cells; this binding then suppresses tip cell fate and expression of Dll4. Notch also regulates VEGFR-2, -3, and the VEGF co-receptor neuropilin 1 (NRP1) and thereby promotes tip/stalk differentiation responses to VEGF and other signals. Notch also decreases cellular migration and enhances pericyte recruitment, consequently helping to stabilize the vessel. HIF1α interacts with the Notch1 intracellular domain and increases Notch target gene expression [121] and, interestingly, restoring the expression of Dll4 rescues most of the effects of deleting HIF2α in ECs in vivo, underscoring the importance of Dll4 as a target of HIF2α [65].
As the branch of the forming vessel extends, the lumen of the new vessel is formed between adjacent endothelial stalk cells [67,122], which allows the flow of oxygenated blood to the tissue via the newly formed vessel. This reduces hypoxia and creates a negative feedback mechanism that then ensures appropriate vascularization. Proper lumen formation requires MMPs and cell–ECM communication, and while it can be accompanied by activation of genes that promote tube regression [123,124], primitive tubes also recruit pericytes and are sequentially covered by an intermediate and a mature ECM, which inhibit regression, stabilize the vessel, and terminate angiogenesis [58,62,125]. Significantly, hypoxia-controlled release of VEGF, ANG1, and ANG2 are some of the main players that regulate pericyte recruitment [126,127,128]. In contrast to the above, exposing brain endothelial cells to several hours of hypoxia disrupts the blood–brain barrier [129], and pericytes or astrocytes cultured under normoxia promote blood–brain barrier stabilization [130,131]. Thus, the precise molecular dance regulating the termination of angiogenesis is still not well understood and the exact roles that HIF1/2 play are still under investigation.
Signals initiating angiogenesis can also come from multiple cell types, including cells from the immune system, such as macrophages. Specifically, tumor-associated macrophages (TAMs) can induce angiogenesis by secreting VEGF, MMPs, and Wnts. Most pro-angiogenic TAMs express Tie2, which can also be upregulated by hypoxia, and Tie2 has been shown to upregulate pro-angiogenic genes, suggesting that it may be a marker for pro-angiogenic macrophages. Members of the Wnt protein family also regulate diverse biological processes, including cellular proliferation, survival, differentiation, migration, and apoptosis, and can act as potent angiogenic factors. They transduce cellular signals by binding to Fzd receptors and members of the lipoprotein-related protein 5 or 6 (LRP5/6) family. Besides the known direct effects of Wnt proteins on angiogenesis, they may also induce expression of inflammatory cytokines in ECs, which can further amplify the angiogenic signal [45,132]. Furthermore, using the pathological retinal neovascularization model, we showed that hematopoietic cell HIF2α can control FasL protein levels in the retina, regulating apoptosis of endothelial cells and pathological neovascularization [133].

5. Maturation of the Blood Vessels

After vessel formation and initiation of blood flow, mesenchymal cells are recruited, which differentiate into SMCs. They then proliferate until wall thickness adequate for providing stability to the vessel has been achieved; thus, the number of SMC layers that will be present in the mature vessel is established early during maturation. Apart from recruitment and differentiation of mural cells, maturation of the vessel depends on the generation of a mature ECM and it is these ECM–mural cell interactions that help stabilize the vessel and suppress angiogenesis. These processes are regulated by several factors, including EC subtype, shear stress, and hypoxia [54,109,111,134]. Specifically, in vivo HIF2α deficiency in ECs results in a variety of phenotypes, including increased vessel permeability, aberrant endothelial cell ultrastructure, and pulmonary hypertension. Moreover, these animals exhibit defective tumor angiogenesis due to increased hypoxic stress and tumor cell apoptosis [69], implicating a role for HIF2α in the regulation of vessel maturation.
Further, ECs secrete Platelet Derived Growth Factor B (PDGFB) during the initial stages of vessel maturation, which facilitates recruitment of mural cell precursors through interaction with the PDGF Receptor β (PDGFRβ) [109,111]. PDGFB is upregulated by HIF1α [2,110], and when added to lung arterial cells, it induces proliferation of smooth muscle cells, promotes the Warburg effect, and activates HIF1α [135]. After recruitment, endothelium-associated mural cell precursors secrete ANG1, while ECs secrete Tie2, and ANG1 binds to Tie2 [136] to promote the formation of cell–cell adhesions and mural cell association with the vessel wall [137]. These events stabilize nascent vessels and render them resistant to leakiness. Notably, in the absence of VEGF (and thus, angiogenic activity), ANG2 can antagonize ANG1, causing vessels to destabilize and regress; however, in the presence of VEGF, which is released in response to hypoxia, ANG2 facilitates a type of vessel destabilization that leads to vascular sprouting rather than regression [109,111]. Of note, hypoxia controls ANG1 and ANG2 expression in a time-dependent manner [68], and besides the interaction between ANG1–Tie2 and EphrinB2 with its receptor EphB4, mural cell precursors also need endothelial contact to adequately differentiate, which requires communication between gap–junction proteins α1 (GJα1) and γ1 (GJγ1), and signal activation via transforming growth factor β1 (TGFβ1) [109,111]. Hypoxia is involved in this process as it regulates the expression of both TGFβ [98,99] and EphrinB2 [89].

6. The Role of Inflammation in Angiogenesis

As mentioned above, the hypoxia pathway is an essential regulator of angiogenesis during development and in the adult. Therefore, it is not surprising that this pathway also plays an important role during stress situations such as inflammation, and hypoxia pathway proteins have been tested as therapeutic targets for reducing inflammation [138,139]. Indeed, HIFs are involved in crucial immune cell features, including survival and glycolysis in neutrophils, and promoting inflammatory functions of certain innate immune cells such as dendritic cells, mast cells, and macrophages [140]. Furthermore, we have recently described a central role for HIF2α in the migration of neutrophils through very confined microenvironments, both ex vivo and under various inflammatory settings in vivo [141]. Conversely, HIFs suppress the adaptive immune response by promoting the differentiation of regulatory T-cells and negatively regulating CD4+ and CD8+ T-cells [139]. Correspondingly, strong cross-communication between endothelial and immune cells during inflammation has also been reported with certain immune cells, e.g., neutrophils and monocytes, reported to play essential roles in inflammation-associated angiogenesis; they also constitute important sources of pro-angiogenic factors such as VEGFA, FGF2, and MMP9, and STAT3-mediated signaling, especially during tumor development [142,143]. Additionally, pro-angiogenic properties of mast cells, i.e., through release of multiple factors such as TNF, FGF2, and VEGFA, have also been reported in certain cancers, such as skin and intestinal adenoma [144]. In vitro, activation of eosinophils leads to secretion of pro-angiogenic factors through degranulation [145], and lymphocytes have also been shown to regulate angiogenesis, both directly and indirectly, such as with the release of similar angiogenic factors by B cells, as by myeloid cells [146]. Similarly, regulatory T-cells promote angiogenesis by releasing VEGFA and thereby suppressing IFNγ-expressing effector TH1 cells. Additionally, while T helper 2 (TH2) cells promote differentiation of tumor-infiltrating monocytes and macrophages into pro-angiogenic tumor-associated macrophages (TAMs), Th1 and Cytotoxic T cells secrete interferon-γ (IFNγ), which inhibits tumor angiogenesis by restricting the proliferation of endothelial cells [147].
Apart from this cross-communication, inflammation and angiogenesis also converge at hypoxia as they share common mediators. Interestingly, several such genes, namely, IL-1β, COX2, and SDF-1, which are regulated by HIFs, are involved in angiogenesis and their roles in the inflammatory response have been described [148,149,150]. Available data show that hypoxia-mediated regulation of angiogenic genes associated with inflammation is predominantly controlled via the NF-κB pathway. For instance, IL-6, COX-2, tumor necrosis factor alpha (TNF-α), macrophage inflammatory protein 2 (MIP-2), intercellular adhesion molecule (ICAM), IL-1β, and inducible NOS (iNOS) are some pro-angiogenic factors that are also secreted by immune cells upon hypoxia-induced NF-κB signaling [151]. Another common factor involved in both angiogenesis and inflammation is nitric oxide (NO), which stabilizes HIF1α during inflammation, leading to induction of pro-angiogenic factors [152,153]. Reports also suggest that the contribution of NO to direct activation and gene transcription of VEGF may be sourced from SMCs in this particular scenario [154,155]. Additionally, NO production in ECs is tightly controlled by HIF1/2α-dependent iNOS expression, and as NO directly regulates vessel permeability, it represents an important regulator of not only immune cell extravasation during inflammatory response but also tumor cell migration through the endothelial layers during metastatic processes [156]. Thus, as the role of hypoxia in angiogenesis and inflammation is remarkable, it represents a viable therapeutic target during inflammatory processes such as tumor development.

7. Hypoxia in the Angiogenic Tumor Environment

Hypoxia is also a characteristic feature of tumor development because tumor vasculature differs from that seen in healthy tissues in that (1) tumor-derived ECs display higher levels of glycolysis and VEGF secretion than normal ECs [27], (2) tumor vessels have a smaller or more compressed lumen, (3) the vessel wall is abnormal with loosely attached mural cells, (4) the ECM is irregular, and (5) the ECs are poorly connected. All of these lead to irregular vascular perfusion that not only affects oxygen and drug delivery but also eventually facilitates metastasis [124].
Research into tumor-associated vasculature has opened at least two possible approaches for therapy, namely, (1) starving the tumor using anti-angiogenic treatment and (2) normalizing tumor vasculature. It has been hypothesized that starving the tumor would arrest its growth, render it dormant, and reduce its size. However, the benefits of restricting angiogenesis have remained modest as such pruning of vessels often leads to increased intra-tumoral hypoxia and consequent radio- and chemo-resistance that trigger pathological angiogenesis and inflammation. Given these disadvantages, another approach, viz., normalizing tumor vasculature, has been proposed. This entails maintaining a balance between pro- and anti-angiogenic factors to reduce vascular permeability and improve blood flow and tumor perfusion [157]. As angiogenesis is strongly linked to hypoxia, it may be possible to target hypoxia pathway proteins to establish this delicate balance between pro- and anti-angiogenic factors. In support of this approach, reports show that in vivo endothelial HIF1α or HIF2α deficiency resulted in enhanced tumor necrosis due to lack of nutrients with consequent reduction in tumor growth and number of tumor vessels [66,69]; in contrast, heterozygous deficiency of PHD2 restored tumor oxygenation and endothelial normalization, inhibiting metastasis [158].

8. Targeting HIF-Mediated Angiogenesis: A Pharmacological Approach

HIF inhibition has emerged as an attractive therapeutic option against cancer that can disrupt the pathological angiogenic switch prompted by tumor cells. Mechanistically, it can involve two major approaches: (1) direct HIF inhibition to sabotage HIF function and stability and, (2) indirect HIF inhibitors that can obstruct upstream or downstream transcriptional capacities [159].
Acriflavine (ACF), a direct HIF inhibitor that prevents HIF heterodimerization, has been shown to inhibit tumor growth and vascularization [160], particularly in glioblastoma, as recent studies on the mechanism of brain tumor growth have revealed a clear association between the overexpression of hypoxia-induced genes and greater probability of invasion, higher recurrence, and poorer clinical outcomes. Mangraviti and colleagues (2017) have evaluated the efficacy of ACF against malignant brain cancer using biodegradable polymers for local sustained drug delivery through the blood–brain barrier. Thus, while in vitro ACF treatment reduced hypoxia-induced overexpression of PGK-1 and VEGF in glioma cell lines, in vivo treatments with different ACF/polymer ratios in a preclinical model led to high survival rates in rats with gliosarcoma [161]. Echinomycin, another well-known direct HIF inhibitor, hampers the binding of HIF1α to the HRE sequence in the VEGF promoter [162], and a recent study by Thomas and collaborators (2005) has demonstrated that nano-delivery of echinomycin induces autophagy-mediated death in pancreatic cancer in vivo, wherein syndecan-1-encapsulated echinomycin, a custom-designed tumor-specific delivery method, resulted in significantly greater survival [163].
Indirect HIF inhibition has also shown promise in recent years and the inhibition of heat shock proteins (particularly Hsp90 and Hsp70) has been reported to have a significant impact on the accumulation of HIF1α proteins in the cytosol, which then promotes their activity. These Hsp chaperones bind to HIF1α and blockade VHL-independent proteasome degradation, permitting proper conformational architecture of HIF1α heterodimers and its subsequent interaction with downstream proteins [164]. The Hsp70 inhibitor IDF-11774, approved for clinical trials phase I study by the Korean Food and Drug Administration, has shown significant dose-dependent reduction in VEGF expression in vivo, along with angiogenesis disruption [165]. More recently, AT-533, a novel Hsp90 inhibitor, has been reported to directly impair HIF1α/VEGF/VEGFR-2-mediated angiogenesis in vitro and in breast carcinoma xenografts in vivo [166]. Hsp inhibitors also severely disrupt tumor-associated angiogenesis, presenting potential benefits for targeted cancer therapy. Additionally, a very promising antiangiogenic and antitumor drug, 2-Methoxyestradiol (2ME2, Panzem®) NanoCrystal® Dispersion (NCD®) can downregulate HIF1α expression by depolarizing cytoplasmatic microtubules that ultimately block HIF1α translation [159,167]. A combination of 2ME2 and bevacizumab (anti-VEGF) has also led to significant tumor reduction in patients with prostate cancer [168].
New molecules that interact with the HIF pathway to disrupt tumor-associated angiogenesis have also been described. For example, Salinomycin, an antimicrobial agent, has shown anticancer potential [169] and more recently, Dewangan and collaborators (2019) have reported that salinomycin suppresses angiogenesis in the 4T1-induced breast carcinoma model [170]. Another study has revealed that SN-38, an active irinotecan metabolite, can inhibit HIF1α expression upon radiation therapy, wherein this inhibition was followed by a reduction in VEGF expression in colorectal cancer cells (HCT116, SW480). Despite such indirect inhibition of HIF1α, the authors also acknowledge that SN-38-induced radiation sensitivity could be achieved by multiple other pathways [171]; nevertheless, these results imply a potential reduction in the harmful effects of increased HIF1α due to radiation therapy. Detailed aspects of clinical development of HIF2α inhibitors, in direct conjunction with VEGF to treat renal cell carcinoma, have been extensively reviewed recently [172]. Although, HIF inhibition seems to be advantageous in cancer therapy, it is important to note that, to date, there are no clinically approved HIF inhibitors. The vital significance of the HIF pathway, accompanied by a nonspecific delivery mechanism, may permit the cautious use of HIF inhibitors.
Conversely, a majority of tumor types overexpress PHDs, making these oxygen sensors attractive therapeutic targets. For a more elaborated overview on the potential impact of PHD inhibitors for the treatment of cancer we refer to a recent review from our research group [173].

9. Conclusions

This review addressed current advances in the biology of hypoxia pathway proteins and their association with components of the vasculature. We have also provided an overview of the development and functional role of the various components and the impact of HIFs. Further understanding of the role of the hypoxia pathway proteins under physiological and pathological settings, in the context of the vasculature, are warranted, and will be of utmost importance in the development of targeted therapies against disorders involving vascular pathologies. In this respect, recent research advocates for the use of combination therapies to target individual insults generated in clinically relevant settings. Nevertheless, additional research is needed to expand our knowledge of the complex mechanisms underlying the effects of hypoxia pathway proteins, namely, HIFs, PHDs, and related genes.

Author Contributions

D.R., D.W., D.G. wrote the manuscript; S.S. and B.W. contributed to the discussion and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the DFG (German Research Foundation) within the CRC/Transregio 205/2, project No. 314061271—TRR205, “The Adrenal: Central Relay in Health and Disease” (A02) to B.W.; DFG grants WI3291/5-1 and 12-1 to B.W. This work was also supported by the DFG priority program µBONE 2084 –T.P 21 (to B.W.). B.W. was further supported by the DFG Heisenberg program (WI3291/13-1). Open Access Funding by the Publication Fund of the TU Dresden.

Informed Consent Statement

Not applicable.

Acknowledgments

We would like to thank Vasuprada Iyengar for English Language and content editing. All figures were created by the authors with BioRender.com.

Conflicts of Interest

The authors declare that no relevant conflict of interest exist.

References

  1. Downes, N.L.; Laham-Karam, N.; Kaikkonen, M.U.; Yla-Herttuala, S. Differential but Complementary HIF1alpha and HIF2alpha Transcriptional Regulation. Mol. Ther. 2018, 26, 1735–1745. [Google Scholar] [CrossRef] [Green Version]
  2. Manalo, D.J.; Rowan, A.; Lavoie, T.; Natarajan, L.; Kelly, B.D.; Ye, S.Q.; Garcia, J.G.; Semenza, G.L. Transcriptional regulation of vascular endothelial cell responses to hypoxia by HIF-1. Blood 2005, 105, 659–669. [Google Scholar] [CrossRef]
  3. Franke, K.; Gassmann, M.; Wielockx, B. Erythrocytosis: The HIF pathway in control. Blood 2013, 122, 1122–1128. [Google Scholar] [CrossRef] [Green Version]
  4. Ivan, M.; Kaelin, W.G., Jr. The EGLN-HIF O2-Sensing System: Multiple Inputs and Feedbacks. Mol. Cell 2017, 66, 772–779. [Google Scholar] [CrossRef] [Green Version]
  5. Wielockx, B.; Grinenko, T.; Mirtschink, P.; Chavakis, T. Hypoxia Pathway Proteins in Normal and Malignant Hematopoiesis. Cells 2019, 8, 155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Meneses, A.M.; Wielockx, B. PHD2: From hypoxia regulation to disease progression. Hypoxia 2016, 4, 53–67. [Google Scholar] [CrossRef] [Green Version]
  7. Benita, Y.; Kikuchi, H.; Smith, A.D.; Zhang, M.Q.; Chung, D.C.; Xavier, R.J. An integrative genomics approach identifies Hypoxia Inducible Factor-1 (HIF-1)-target genes that form the core response to hypoxia. Nucleic Acids Res. 2009, 37, 4587–4602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Koivunen, P.; Hirsila, M.; Gunzler, V.; Kivirikko, K.I.; Myllyharju, J. Catalytic properties of the asparaginyl hydroxylase (FIH) in the oxygen sensing pathway are distinct from those of its prolyl 4-hydroxylases. J. Biol. Chem. 2004, 279, 9899–9904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Foxler, D.E.; Bridge, K.S.; Foster, J.G.; Grevitt, P.; Curry, S.; Shah, K.M.; Davidson, K.M.; Nagano, A.; Gadaleta, E.; Rhys, H.I.; et al. A HIF-LIMD1 negative feedback mechanism mitigates the pro-tumorigenic effects of hypoxia. EMBO Mol. Med. 2018, 10, e8340. [Google Scholar] [CrossRef] [PubMed]
  10. Foxler, D.E.; Bridge, K.S.; James, V.; Webb, T.M.; Mee, M.; Wong, S.C.; Feng, Y.; Constantin-Teodosiu, D.; Petursdottir, T.E.; Bjornsson, J.; et al. The LIMD1 protein bridges an association between the prolyl hydroxylases and VHL to repress HIF-1 activity. Nat. Cell Biol. 2012, 14, 201–208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Park, M.H.; Choi, K.Y.; Jung, Y.; Min do, S. Phospholipase D1 protein coordinates dynamic assembly of HIF-1alpha-PHD-VHL to regulate HIF-1alpha stability. Oncotarget 2014, 5, 11857–11872. [Google Scholar] [CrossRef] [Green Version]
  12. Wong, W.; Goehring, A.S.; Kapiloff, M.S.; Langeberg, L.K.; Scott, J.D. mAKAP compartmentalizes oxygen-dependent control of HIF-1alpha. Sci. Signal. 2008, 1, ra18. [Google Scholar] [CrossRef] [Green Version]
  13. Choi, Y.K.; Kim, J.H.; Kim, W.J.; Lee, H.Y.; Park, J.A.; Lee, S.W.; Yoon, D.K.; Kim, H.H.; Chung, H.; Yu, Y.S.; et al. AKAP12 regulates human blood-retinal barrier formation by downregulation of hypoxia-inducible factor-1alpha. J. Neurosci. 2007, 27, 4472–4481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Lee, S.H.; Bae, S.C.; Kim, K.W.; Lee, Y.M. RUNX3 inhibits hypoxia-inducible factor-1alpha protein stability by interacting with prolyl hydroxylases in gastric cancer cells. Oncogene 2014, 33, 1458–1467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Colla, S.; Tagliaferri, S.; Morandi, F.; Lunghi, P.; Donofrio, G.; Martorana, D.; Mancini, C.; Lazzaretti, M.; Mazzera, L.; Ravanetti, L.; et al. The new tumor-suppressor gene inhibitor of growth family member 4 (ING4) regulates the production of proangiogenic molecules by myeloma cells and suppresses hypoxia-inducible factor-1 alpha (HIF-1alpha) activity: Involvement in myeloma-induced angiogenesis. Blood 2007, 110, 4464–4475. [Google Scholar] [CrossRef] [PubMed]
  16. Song, D.; Li, L.S.; Heaton-Johnson, K.J.; Arsenault, P.R.; Master, S.R.; Lee, F.S. Prolyl hydroxylase domain protein 2 (PHD2) binds a Pro-Xaa-Leu-Glu motif, linking it to the heat shock protein 90 pathway. J. Biol. Chem. 2013, 288, 9662–9674. [Google Scholar] [CrossRef] [Green Version]
  17. Baek, J.H.; Mahon, P.C.; Oh, J.; Kelly, B.; Krishnamachary, B.; Pearson, M.; Chan, D.A.; Giaccia, A.J.; Semenza, G.L. OS-9 interacts with hypoxia-inducible factor 1alpha and prolyl hydroxylases to promote oxygen-dependent degradation of HIF-1alpha. Mol. Cell 2005, 17, 503–512. [Google Scholar] [CrossRef]
  18. Melillo, G.; Musso, T.; Sica, A.; Taylor, L.S.; Cox, G.W.; Varesio, L. A hypoxia-responsive element mediates a novel pathway of activation of the inducible nitric oxide synthase promoter. J. Exp. Med. 1995, 182, 1683–1693. [Google Scholar] [CrossRef] [Green Version]
  19. Dehne, N.; Brune, B. HIF-1 in the inflammatory microenvironment. Exp. Cell Res. 2009, 315, 1791–1797. [Google Scholar] [CrossRef]
  20. Yin, J.H.; Yang, D.I.; Ku, G.; Hsu, C.Y. iNOS expression inhibits hypoxia-inducible factor-1 activity. Biochem. Biophys. Res. Commun. 2000, 279, 30–34. [Google Scholar] [CrossRef]
  21. Keith, B.; Johnson, R.S.; Simon, M.C. HIF1alpha and HIF2alpha: Sibling rivalry in hypoxic tumour growth and progression. Nat. Rev. Cancer 2011, 12, 9–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Kuo, T.C.; Kung, H.J.; Shih, J.W. Signaling in and out: Long-noncoding RNAs in tumor hypoxia. J. Biomed. Sci. 2020, 27, 59. [Google Scholar] [CrossRef] [PubMed]
  23. Nauta, T.D.; van den Broek, M.; Gibbs, S.; van der Pouw-Kraan, T.C.; Oudejans, C.B.; van Hinsbergh, V.W.; Koolwijk, P. Identification of HIF-2alpha-regulated genes that play a role in human microvascular endothelial sprouting during prolonged hypoxia in vitro. Angiogenesis 2017, 20, 39–54. [Google Scholar] [CrossRef] [Green Version]
  24. Weigand, J.E.; Boeckel, J.N.; Gellert, P.; Dimmeler, S. Hypoxia-induced alternative splicing in endothelial cells. PLoS ONE 2012, 7, e42697. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Lee, D.C.; Sohn, H.A.; Park, Z.Y.; Oh, S.; Kang, Y.K.; Lee, K.M.; Kang, M.; Jang, Y.J.; Yang, S.J.; Hong, Y.K.; et al. A lactate-induced response to hypoxia. Cell 2015, 161, 595–609. [Google Scholar] [CrossRef] [Green Version]
  26. Ruan, G.X.; Kazlauskas, A. Lactate engages receptor tyrosine kinases Axl, Tie2, and vascular endothelial growth factor receptor 2 to activate phosphoinositide 3-kinase/Akt and promote angiogenesis. J. Biol. Chem. 2013, 288, 21161–21172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Wong, B.W.; Marsch, E.; Treps, L.; Baes, M.; Carmeliet, P. Endothelial cell metabolism in health and disease: Impact of hypoxia. EMBO J. 2017, 36, 2187–2203. [Google Scholar] [CrossRef]
  28. Caprioli, A.; Minko, K.; Drevon, C.; Eichmann, A.; Dieterlen-Lievre, F.; Jaffredo, T. Hemangioblast commitment in the avian allantois: Cellular and molecular aspects. Dev. Biol. 2001, 238, 64–78. [Google Scholar] [CrossRef] [Green Version]
  29. Ribatti, D.; Nico, B.; Crivellato, E. Morphological and molecular aspects of physiological vascular morphogenesis. Angiogenesis 2009, 12, 101–111. [Google Scholar] [CrossRef]
  30. Risau, W. Mechanisms of angiogenesis. Nature 1997, 386, 671–674. [Google Scholar] [CrossRef]
  31. Cox, C.M.; Poole, T.J. Angioblast differentiation is influenced by the local environment: FGF-2 induces angioblasts and patterns vessel formation in the quail embryo. Dev. Dyn. 2000, 218, 371–382. [Google Scholar] [CrossRef]
  32. Kakudo, N.; Morimoto, N.; Ogawa, T.; Taketani, S.; Kusumoto, K. Hypoxia Enhances Proliferation of Human Adipose-Derived Stem Cells via HIF-1a Activation. PLoS ONE 2015, 10, e0139890. [Google Scholar] [CrossRef] [Green Version]
  33. Rosmorduc, O.; Wendum, D.; Corpechot, C.; Galy, B.; Sebbagh, N.; Raleigh, J.; Housset, C.; Poupon, R. Hepatocellular hypoxia-induced vascular endothelial growth factor expression and angiogenesis in experimental biliary cirrhosis. Am. J. Pathol. 1999, 155, 1065–1073. [Google Scholar] [CrossRef] [Green Version]
  34. Conte, C.; Riant, E.; Toutain, C.; Pujol, F.; Arnal, J.F.; Lenfant, F.; Prats, A.C. FGF2 translationally induced by hypoxia is involved in negative and positive feedback loops with HIF-1alpha. PLoS ONE 2008, 3, e3078. [Google Scholar] [CrossRef] [PubMed]
  35. Grimm, C.; Wenzel, A.; Groszer, M.; Mayser, H.; Seeliger, M.; Samardzija, M.; Bauer, C.; Gassmann, M.; Reme, C.E. HIF-1-induced erythropoietin in the hypoxic retina protects against light-induced retinal degeneration. Nat. Med. 2002, 8, 718–724. [Google Scholar] [CrossRef]
  36. Murakami, M.; Nguyen, L.T.; Hatanaka, K.; Schachterle, W.; Chen, P.Y.; Zhuang, Z.W.; Black, B.L.; Simons, M. FGF-dependent regulation of VEGF receptor 2 expression in mice. J. Clin. Investig. 2011, 121, 2668–2678. [Google Scholar] [CrossRef] [Green Version]
  37. Carmeliet, P. Mechanisms of angiogenesis and arteriogenesis. Nat. Med. 2000, 6, 389–395. [Google Scholar] [CrossRef]
  38. 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] [Green Version]
  39. Gerber, H.P.; Condorelli, F.; Park, J.; Ferrara, N. Differential transcriptional regulation of the two vascular endothelial growth factor receptor genes. Flt-1, but not Flk-1/KDR, is up-regulated by hypoxia. J. Biol. Chem. 1997, 272, 23659–23667. [Google Scholar] [CrossRef] [Green Version]
  40. Tscheudschilsuren, G.; Aust, G.; Nieber, K.; Schilling, N.; Spanel-Borowski, K. Microvascular endothelial cells differ in basal and hypoxia-regulated expression of angiogenic factors and their receptors. Microvasc. Res. 2002, 63, 243–251. [Google Scholar] [CrossRef]
  41. Iyer, N.V.; Kotch, L.E.; Agani, F.; Leung, S.W.; Laughner, E.; Wenger, R.H.; Gassmann, M.; Gearhart, J.D.; Lawler, A.M.; Yu, A.Y.; et al. Cellular and developmental control of O2 homeostasis by hypoxia-inducible factor 1 alpha. Genes Dev. 1998, 12, 149–162. [Google Scholar] [CrossRef] [Green Version]
  42. Imanirad, P.; Dzierzak, E. Hypoxia and HIFs in regulating the development of the hematopoietic system. Blood Cells Mol. Dis. 2013, 51, 256–263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Ramirez-Bergeron, D.L.; Runge, A.; Adelman, D.M.; Gohil, M.; Simon, M.C. HIF-dependent hematopoietic factors regulate the development of the embryonic vasculature. Dev. Cell 2006, 11, 81–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Rae, F.; Woods, K.; Sasmono, T.; Campanale, N.; Taylor, D.; Ovchinnikov, D.A.; Grimmond, S.M.; Hume, D.A.; Ricardo, S.D.; Little, M.H. Characterisation and trophic functions of murine embryonic macrophages based upon the use of a Csf1r-EGFP transgene reporter. Dev. Biol. 2007, 308, 232–246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Newman, A.C.; Hughes, C.C. Macrophages and angiogenesis: A role for Wnt signaling. Vasc. Cell 2012, 4, 13. [Google Scholar] [CrossRef] [Green Version]
  46. Gerri, C.; Marin-Juez, R.; Marass, M.; Marks, A.; Maischein, H.M.; Stainier, D.Y.R. Hif-1alpha regulates macrophage-endothelial interactions during blood vessel development in zebrafish. Nat. Commun. 2017, 8, 15492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Scortegagna, M.; Morris, M.A.; Oktay, Y.; Bennett, M.; Garcia, J.A. The HIF family member EPAS1/HIF-2alpha is required for normal hematopoiesis in mice. Blood 2003, 102, 1634–1640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Rankin, E.B.; Rha, J.; Unger, T.L.; Wu, C.H.; Shutt, H.P.; Johnson, R.S.; Simon, M.C.; Keith, B.; Haase, V.H. Hypoxia-inducible factor-2 regulates vascular tumorigenesis in mice. Oncogene 2008, 27, 5354–5358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Rankin, E.B.; Biju, M.P.; Liu, Q.; Unger, T.L.; Rha, J.; Johnson, R.S.; Simon, M.C.; Keith, B.; Haase, V.H. Hypoxia-inducible factor-2 (HIF-2) regulates hepatic erythropoietin in vivo. J. Clin. Investig. 2007, 117, 1068–1077. [Google Scholar] [CrossRef]
  50. Yoon, D.; Pastore, Y.D.; Divoky, V.; Liu, E.; Mlodnicka, A.E.; Rainey, K.; Ponka, P.; Semenza, G.L.; Schumacher, A.; Prchal, J.T. Hypoxia-inducible factor-1 deficiency results in dysregulated erythropoiesis signaling and iron homeostasis in mouse development. J. Biol. Chem. 2006, 281, 25703–25711. [Google Scholar] [CrossRef] [Green Version]
  51. Imanirad, P.; Solaimani Kartalaei, P.; Crisan, M.; Vink, C.; Yamada-Inagawa, T.; de Pater, E.; Kurek, D.; Kaimakis, P.; van der Linden, R.; Speck, N.; et al. HIF1alpha is a regulator of hematopoietic progenitor and stem cell development in hypoxic sites of the mouse embryo. Stem Cell Res. 2014, 12, 24–35. [Google Scholar] [CrossRef] [Green Version]
  52. Cowden Dahl, K.D.; Fryer, B.H.; Mack, F.A.; Compernolle, V.; Maltepe, E.; Adelman, D.M.; Carmeliet, P.; Simon, M.C. Hypoxia-inducible factors 1alpha and 2alpha regulate trophoblast differentiation. Mol. Cell. Biol. 2005, 25, 10479–10491. [Google Scholar] [CrossRef] [Green Version]
  53. Schaffer, L.; Vogel, J.; Breymann, C.; Gassmann, M.; Marti, H.H. Preserved placental oxygenation and development during severe systemic hypoxia. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2006, 290, R844–R851. [Google Scholar] [CrossRef] [PubMed]
  54. Germain, S.; Monnot, C.; Muller, L.; Eichmann, A. Hypoxia-driven angiogenesis: Role of tip cells and extracellular matrix scaffolding. Curr. Opin. Hematol. 2010, 17, 245–251. [Google Scholar] [CrossRef]
  55. Petrova, V.; Annicchiarico-Petruzzelli, M.; Melino, G.; Amelio, I. The hypoxic tumour microenvironment. Oncogenesis 2018, 7, 10. [Google Scholar] [CrossRef] [PubMed]
  56. Wagenseil, J.E.; Mecham, R.P. Vascular extracellular matrix and arterial mechanics. Physiol. Rev. 2009, 89, 957–989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Das, S.; Kumar, K.N. Nitric oxide: Its identity and role in blood pressure control. Life Sci. 1995, 57, 1547–1556. [Google Scholar] [CrossRef]
  58. Rhodes, J.M.; Simons, M. The extracellular matrix and blood vessel formation: Not just a scaffold. J. Cell Mol. Med. 2007, 11, 176–205. [Google Scholar] [CrossRef] [PubMed]
  59. Takada, Y.; Ye, X.; Simon, S. The integrins. Genome Biol. 2007, 8, 215. [Google Scholar] [CrossRef] [Green Version]
  60. Kiouptsi, K.; Gambaryan, S.; Walter, E.; Walter, U.; Jurk, K.; Reinhardt, C. Hypoxia impairs agonist-induced integrin alphaIIbbeta3 activation and platelet aggregation. Sci. Rep. 2017, 7, 7621. [Google Scholar] [CrossRef]
  61. Watts, D.; Gaete, D.; Rodriguez, D.; Hoogewijs, D.; Rauner, M.; Sormendi, S.; Wielockx, B. Hypoxia Pathway Proteins are Master Regulators of Erythropoiesis. Int. J. Mol. Sci. 2020, 21, 8131. [Google Scholar] [CrossRef]
  62. Senger, D.R.; Davis, G.E. Angiogenesis. Cold Spring Harb. Perspect. Biol. 2011, 3, a005090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Eichmann, A.; Makinen, T.; Alitalo, K. Neural guidance molecules regulate vascular remodeling and vessel navigation. Genes Dev. 2005, 19, 1013–1021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Sarabipour, S.; Mac Gabhann, F. VEGF-A121a binding to Neuropilins—A concept revisited. Cell Adhes. Migr. 2018, 12, 204–214. [Google Scholar] [CrossRef]
  65. Skuli, N.; Majmundar, A.J.; Krock, B.L.; Mesquita, R.C.; Mathew, L.K.; Quinn, Z.L.; Runge, A.; Liu, L.; Kim, M.N.; Liang, J.; et al. Endothelial HIF-2alpha regulates murine pathological angiogenesis and revascularization processes. J. Clin. Investig. 2012, 122, 1427–1443. [Google Scholar] [CrossRef]
  66. Tang, N.; Wang, L.; Esko, J.; Giordano, F.J.; Huang, Y.; Gerber, H.P.; Ferrara, N.; Johnson, R.S. Loss of HIF-1alpha in endothelial cells disrupts a hypoxia-driven VEGF autocrine loop necessary for tumorigenesis. Cancer Cell 2004, 6, 485–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Kim, J.; Kim, Y.H.; Kim, J.; Park, D.Y.; Bae, H.; Lee, D.H.; Kim, K.H.; Hong, S.P.; Jang, S.P.; Kubota, Y.; et al. YAP/TAZ regulates sprouting angiogenesis and vascular barrier maturation. J. Clin. Investig. 2017, 127, 3441–3461. [Google Scholar] [CrossRef]
  68. Ray, P.S.; Estrada-Hernandez, T.; Sasaki, H.; Zhu, L.; Maulik, N. Early effects of hypoxia/reoxygenation on VEGF, ang-1, ang-2 and their receptors in the rat myocardium: Implications for myocardial angiogenesis. Mol. Cell. Biochem. 2000, 213, 145–153. [Google Scholar] [CrossRef]
  69. Skuli, N.; Liu, L.; Runge, A.; Wang, T.; Yuan, L.; Patel, S.; Iruela-Arispe, L.; Simon, M.C.; Keith, B. Endothelial deletion of hypoxia-inducible factor-2alpha (HIF-2alpha) alters vascular function and tumor angiogenesis. Blood 2009, 114, 469–477. [Google Scholar] [CrossRef] [Green Version]
  70. 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]
  71. Enholm, B.; Paavonen, K.; Ristimaki, A.; Kumar, V.; Gunji, Y.; Klefstrom, J.; Kivinen, L.; Laiho, M.; Olofsson, B.; Joukov, V.; et al. Comparison of VEGF, VEGF-B, VEGF-C and Ang-1 mRNA regulation by serum, growth factors, oncoproteins and hypoxia. Oncogene 1997, 14, 2475–2483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Varela-Nallar, L.; Rojas-Abalos, M.; Abbott, A.C.; Moya, E.A.; Iturriaga, R.; Inestrosa, N.C. Chronic hypoxia induces the activation of the Wnt/beta-catenin signaling pathway and stimulates hippocampal neurogenesis in wild-type and APPswe-PS1DeltaE9 transgenic mice in vivo. Front. Cell. Neurosci. 2014, 8, 17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Lim, J.H.; Chun, Y.S.; Park, J.W. Hypoxia-inducible factor-1alpha obstructs a Wnt signaling pathway by inhibiting the hARD1-mediated activation of beta-catenin. Cancer Res. 2008, 68, 5177–5184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Verras, M.; Papandreou, I.; Lim, A.L.; Denko, N.C. Tumor hypoxia blocks Wnt processing and secretion through the induction of endoplasmic reticulum stress. Mol. Cell. Biol. 2008, 28, 7212–7224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Qi, C.; Zhang, J.; Chen, X.; Wan, J.; Wang, J.; Zhang, P.; Liu, Y. Hypoxia stimulates neural stem cell proliferation by increasing HIF1alpha expression and activating Wnt/beta-catenin signaling. Cell. Mol. Biol. Noisy-Le-Grand 2017, 63, 12–19. [Google Scholar] [CrossRef] [PubMed]
  76. Yuan, K.; Orcholski, M.E.; Panaroni, C.; Shuffle, E.M.; Huang, N.F.; Jiang, X.; Tian, W.; Vladar, E.K.; Wang, L.; Nicolls, M.R.; et al. Activation of the Wnt/planar cell polarity pathway is required for pericyte recruitment during pulmonary angiogenesis. Am. J. Pathol. 2015, 185, 69–84. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Zerlin, M.; Julius, M.A.; Kitajewski, J. Wnt/Frizzled signaling in angiogenesis. Angiogenesis 2008, 11, 63–69. [Google Scholar] [CrossRef]
  78. Ramakrishnan, S.; Anand, V.; Roy, S. Vascular endothelial growth factor signaling in hypoxia and inflammation. J. Neuroimmune Pharmacol. 2014, 9, 142–160. [Google Scholar] [CrossRef] [Green Version]
  79. Pugh, C.W.; Ratcliffe, P.J. Regulation of angiogenesis by hypoxia: Role of the HIF system. Nat. Med. 2003, 9, 677–684. [Google Scholar] [CrossRef]
  80. Bartoszewska, S.; Kochan, K.; Piotrowski, A.; Kamysz, W.; Ochocka, R.J.; Collawn, J.F.; Bartoszewski, R. The hypoxia-inducible miR-429 regulates hypoxia-inducible factor-1alpha expression in human endothelial cells through a negative feedback loop. FASEB J. 2015, 29, 1467–1479. [Google Scholar] [CrossRef] [Green Version]
  81. 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]
  82. Chen, W.G.; Sun, J.; Shen, W.W.; Yang, S.Z.; Zhang, Y.; Hu, X.; Qiu, H.; Xu, S.C.; Chu, T.W. Sema4D expression and secretion are increased by HIF-1alpha and inhibit osteogenesis in bone metastases of lung cancer. Clin. Exp. Metastasis 2019, 36, 39–56. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, X.; Sun, Y.; Tian, W.; Wang, F.; Lv, X.; Wang, M.; Sun, T.; Zhang, J.; Wang, L.; Han, M. Sema4A Responds to Hypoxia and Is Involved in Breast Cancer Progression. Biol. Pharm. Bull. 2018, 41, 1791–1796. [Google Scholar] [CrossRef]
  84. Sun, Q.; Zhou, H.; Binmadi, N.O.; Basile, J.R. Hypoxia-inducible factor-1-mediated regulation of semaphorin 4D affects tumor growth and vascularity. J. Biol. Chem. 2009, 284, 32066–32074. [Google Scholar] [CrossRef] [Green Version]
  85. Coma, S.; Shimizu, A.; Klagsbrun, M. Hypoxia induces tumor and endothelial cell migration in a semaphorin 3F- and VEGF-dependent manner via transcriptional repression of their common receptor neuropilin 2. Cell Adhes. Migr. 2011, 5, 266–275. [Google Scholar] [CrossRef] [Green Version]
  86. Misra, R.M.; Bajaj, M.S.; Kale, V.P. Vasculogenic mimicry of HT1080 tumour cells in vivo: Critical role of HIF-1alpha-neuropilin-1 axis. PLoS ONE 2012, 7, e50153. [Google Scholar] [CrossRef]
  87. Ota, H.; Katsube, K.; Ogawa, J.; Yanagishita, M. Hypoxia/Notch signaling in primary culture of rat lymphatic endothelial cells. FEBS Lett. 2007, 581, 5220–5226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. de Bruin, A.; PW, A.C.; Kirchmaier, B.C.; Mokry, M.; Iich, E.; Nirmala, E.; Liang, K.H.; AM, D.V.; Scholman, K.T.; Groot Koerkamp, M.J.; et al. Genome-wide analysis reveals NRP1 as a direct HIF1alpha-E2F7 target in the regulation of motorneuron guidance in vivo. Nucleic Acids Res. 2016, 44, 3549–3566. [Google Scholar] [CrossRef] [Green Version]
  89. Sohl, M.; Lanner, F.; Farnebo, F. Sp1 mediate hypoxia induced ephrinB2 expression via a hypoxia-inducible factor independent mechanism. Biochem. Biophys. Res. Commun. 2010, 391, 24–27. [Google Scholar] [CrossRef]
  90. Vihanto, M.M.; Plock, J.; Erni, D.; Frey, B.M.; Frey, F.J.; Huynh-Do, U. Hypoxia up-regulates expression of Eph receptors and ephrins in mouse skin. FASEB J. 2005, 19, 1689–1691. [Google Scholar] [CrossRef] [Green Version]
  91. Rosenberger, P.; Schwab, J.M.; Mirakaj, V.; Masekowsky, E.; Mager, A.; Morote-Garcia, J.C.; Unertl, K.; Eltzschig, H.K. Hypoxia-inducible factor-dependent induction of netrin-1 dampens inflammation caused by hypoxia. Nat. Immunol. 2009, 10, 195–202. [Google Scholar] [CrossRef] [PubMed]
  92. Ramkhelawon, B.; Yang, Y.; van Gils, J.M.; Hewing, B.; Rayner, K.J.; Parathath, S.; Guo, L.; Oldebeken, S.; Feig, J.L.; Fisher, E.A.; et al. Hypoxia induces netrin-1 and Unc5b in atherosclerotic plaques: Mechanism for macrophage retention and survival. Arterioscler. Thromb. Vasc. Biol. 2013, 33, 1180–1188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Liao, W.X.; Laurent, L.C.; Agent, S.; Hodges, J.; Chen, D.B. Human placental expression of SLIT/ROBO signaling cues: Effects of preeclampsia and hypoxia. Biol. Reprod. 2012, 86, 111. [Google Scholar] [CrossRef]
  94. Hiyama, A.; Skubutyte, R.; Markova, D.; Anderson, D.G.; Yadla, S.; Sakai, D.; Mochida, J.; Albert, T.J.; Shapiro, I.M.; Risbud, M.V. Hypoxia activates the notch signaling pathway in cells of the intervertebral disc: Implications in degenerative disc disease. Arthritis Rheum. 2011, 63, 1355–1364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Liu, H.; Zhang, W.; Kennard, S.; Caldwell, R.B.; Lilly, B. Notch3 is critical for proper angiogenesis and mural cell investment. Circ. Res. 2010, 107, 860–870. [Google Scholar] [CrossRef]
  96. Yao, L.; Kan, E.M.; Kaur, C.; Dheen, S.T.; Hao, A.; Lu, J.; Ling, E.A. Notch-1 signaling regulates microglia activation via NF-kappaB pathway after hypoxic exposure in vivo and in vitro. PLoS ONE 2013, 8, e78439. [Google Scholar] [CrossRef] [Green Version]
  97. Kiriakidis, S.; Henze, A.T.; Kruszynska-Ziaja, I.; Skobridis, K.; Theodorou, V.; Paleolog, E.M.; Mazzone, M. Factor-inhibiting HIF-1 (FIH-1) is required for human vascular endothelial cell survival. FASEB J. 2015, 29, 2814–2827. [Google Scholar] [CrossRef]
  98. Balamurugan, K. HIF-1 at the crossroads of hypoxia, inflammation, and cancer. Int. J. Cancer 2016, 138, 1058–1066. [Google Scholar] [CrossRef] [PubMed]
  99. Copple, B.L. Hypoxia stimulates hepatocyte epithelial to mesenchymal transition by hypoxia-inducible factor and transforming growth factor-beta-dependent mechanisms. Liver Int. Off. J. Int. Assoc. Study Liver 2010, 30, 669–682. [Google Scholar] [CrossRef] [Green Version]
  100. Wei, H.; Bedja, D.; Koitabashi, N.; Xing, D.; Chen, J.; Fox-Talbot, K.; Rouf, R.; Chen, S.; Steenbergen, C.; Harmon, J.W.; et al. Endothelial expression of hypoxia-inducible factor 1 protects the murine heart and aorta from pressure overload by suppression of TGF-beta signaling. Proc. Natl. Acad. Sci. USA 2012, 109, E841–E850. [Google Scholar] [CrossRef] [Green Version]
  101. 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] [Green Version]
  102. Moreno-Miralles, I.; Ren, R.; Moser, M.; Hartnett, M.E.; Patterson, C. Bone morphogenetic protein endothelial cell precursor-derived regulator regulates retinal angiogenesis in vivo in a mouse model of oxygen-induced retinopathy. Arterioscler. Thromb. Vasc. Biol. 2011, 31, 2216–2222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Tian, F.; Zhou, A.X.; Smits, A.M.; Larsson, E.; Goumans, M.J.; Heldin, C.H.; Boren, J.; Akyurek, L.M. Endothelial cells are activated during hypoxia via endoglin/ALK-1/SMAD1/5 signaling in vivo and in vitro. Biochem. Biophys. Res. Commun. 2010, 392, 283–288. [Google Scholar] [CrossRef] [PubMed]
  104. Zhou, H.; Wu, G.; Ma, X.; Xiao, J.; Yu, G.; Yang, C.; Xu, N.; Zhang, B.; Zhou, J.; Ye, Z.; et al. Attenuation of TGFBR2 expression and tumour progression in prostate cancer involve diverse hypoxia-regulated pathways. J. Exp. Clin. Cancer Res. 2018, 37, 89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Sanchez-Elsner, T.; Botella, L.M.; Velasco, B.; Langa, C.; Bernabeu, C. Endoglin expression is regulated by transcriptional cooperation between the hypoxia and transforming growth factor-beta pathways. J. Biol. Chem. 2002, 277, 43799–43808. [Google Scholar] [CrossRef] [Green Version]
  106. Vicencio, A.G.; Eickelberg, O.; Stankewich, M.C.; Kashgarian, M.; Haddad, G.G. Regulation of TGF-beta ligand and receptor expression in neonatal rat lungs exposed to chronic hypoxia. J. Appl. Physiol. 2002, 93, 1123–1130. [Google Scholar] [CrossRef]
  107. Yoon, H.; Lim, J.H.; Cho, C.H.; Huang, L.E.; Park, J.W. CITED2 controls the hypoxic signaling by snatching p300 from the two distinct activation domains of HIF-1alpha. Biochim. Biophys. Acta 2011, 1813, 2008–2016. [Google Scholar] [CrossRef] [Green Version]
  108. Withington, S.L.; Scott, A.N.; Saunders, D.N.; Lopes Floro, K.; Preis, J.I.; Michalicek, J.; Maclean, K.; Sparrow, D.B.; Barbera, J.P.; Dunwoodie, S.L. Loss of Cited2 affects trophoblast formation and vascularization of the mouse placenta. Dev. Biol. 2006, 294, 67–82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Jain, R.K. Molecular regulation of vessel maturation. Nat. Med. 2003, 9, 685–693. [Google Scholar] [CrossRef] [PubMed]
  110. Yoshida, D.; Kim, K.; Noha, M.; Teramoto, A. Hypoxia inducible factor 1-alpha regulates of platelet derived growth factor-B in human glioblastoma cells. J. Neurooncol. 2006, 76, 13–21. [Google Scholar] [CrossRef] [PubMed]
  111. Barlow, K.D.; Sanders, A.M.; Soker, S.; Ergun, S.; Metheny-Barlow, L.J. Pericytes on the tumor vasculature: Jekyll or hyde? Cancer Microenviron. 2013, 6, 1–17. [Google Scholar] [CrossRef] [Green Version]
  112. Kimura, H.; Miyashita, H.; Suzuki, Y.; Kobayashi, M.; Watanabe, K.; Sonoda, H.; Ohta, H.; Fujiwara, T.; Shimosegawa, T.; Sato, Y. Distinctive localization and opposed roles of vasohibin-1 and vasohibin-2 in the regulation of angiogenesis. Blood 2009, 113, 4810–4818. [Google Scholar] [CrossRef] [Green Version]
  113. Kozako, T.; Matsumoto, N.; Kuramoto, Y.; Sakata, A.; Motonagare, R.; Aikawa, A.; Imoto, M.; Toda, A.; Honda, S.; Shimeno, H.; et al. Vasohibin induces prolyl hydroxylase-mediated degradation of hypoxia-inducible factor-1alpha in human umbilical vein endothelial cells. FEBS Lett. 2012, 586, 1067–1072. [Google Scholar] [CrossRef] [Green Version]
  114. Watanabe, K.; Hasegawa, Y.; Yamashita, H.; Shimizu, K.; Ding, Y.; Abe, M.; Ohta, H.; Imagawa, K.; Hojo, K.; Maki, H.; et al. Vasohibin as an endothelium-derived negative feedback regulator of angiogenesis. J. Clin. Investig. 2004, 114, 898–907. [Google Scholar] [CrossRef] [Green Version]
  115. Augustin, H.G.; Koh, G.Y.; Thurston, G.; Alitalo, K. Control of vascular morphogenesis and homeostasis through the angiopoietin-Tie system. Nat. Rev. Mol. Cell Biol. 2009, 10, 165–177. [Google Scholar] [CrossRef]
  116. Lamalice, L.; Le Boeuf, F.; Huot, J. Endothelial cell migration during angiogenesis. Circ. Res. 2007, 100, 782–794. [Google Scholar] [CrossRef]
  117. De Smet, F.; Segura, I.; De Bock, K.; Hohensinner, P.J.; Carmeliet, P. Mechanisms of vessel branching: Filopodia on endothelial tip cells lead the way. Arterioscler. Thromb. Vasc. Biol. 2009, 29, 639–649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Brantley-Sieders, D.M.; Chen, J. Eph receptor tyrosine kinases in angiogenesis: From development to disease. Angiogenesis 2004, 7, 17–28. [Google Scholar] [CrossRef]
  119. Gerhardt, H.; Golding, M.; Fruttiger, M.; Ruhrberg, C.; Lundkvist, A.; Abramsson, A.; Jeltsch, M.; Mitchell, C.; Alitalo, K.; Shima, D.; et al. VEGF guides angiogenic sprouting utilizing endothelial tip cell filopodia. J. Cell Biol. 2003, 161, 1163–1177. [Google Scholar] [CrossRef] [PubMed]
  120. Tammela, T.; Zarkada, G.; Wallgard, E.; Murtomaki, A.; Suchting, S.; Wirzenius, M.; Waltari, M.; Hellstrom, M.; Schomber, T.; Peltonen, R.; et al. Blocking VEGFR-3 suppresses angiogenic sprouting and vascular network formation. Nature 2008, 454, 656–660. [Google Scholar] [CrossRef] [PubMed]
  121. Carmeliet, P.; De Smet, F.; Loges, S.; Mazzone, M. Branching morphogenesis and antiangiogenesis candidates: Tip cells lead the way. Nat. Rev. Clin. Oncol. 2009, 6, 315–326. [Google Scholar] [CrossRef] [PubMed]
  122. Strilic, B.; Kucera, T.; Eglinger, J.; Hughes, M.R.; McNagny, K.M.; Tsukita, S.; Dejana, E.; Ferrara, N.; Lammert, E. The molecular basis of vascular lumen formation in the developing mouse aorta. Dev. Cell 2009, 17, 505–515. [Google Scholar] [CrossRef] [Green Version]
  123. Iruela-Arispe, M.L.; Davis, G.E. Cellular and molecular mechanisms of vascular lumen formation. Dev. Cell 2009, 16, 222–231. [Google Scholar] [CrossRef] [Green Version]
  124. Potente, M.; Gerhardt, H.; Carmeliet, P. Basic and therapeutic aspects of angiogenesis. Cell 2011, 146, 873–887. [Google Scholar] [CrossRef] [Green Version]
  125. Chambers, R.C.; Leoni, P.; Kaminski, N.; Laurent, G.J.; Heller, R.A. Global expression profiling of fibroblast responses to transforming growth factor-beta1 reveals the induction of inhibitor of differentiation-1 and provides evidence of smooth muscle cell phenotypic switching. Am. J. Pathol. 2003, 162, 533–546. [Google Scholar] [CrossRef]
  126. Cossutta, M.; Darche, M.; Carpentier, G.; Houppe, C.; Ponzo, M.; Raineri, F.; Vallee, B.; Gilles, M.E.; Villain, D.; Picard, E.; et al. Weibel-Palade Bodies Orchestrate Pericytes During Angiogenesis. Arterioscler. Thromb. Vasc. Biol. 2019, 39, 1843–1858. [Google Scholar] [CrossRef] [PubMed]
  127. Park, Y.S.; Kim, N.H.; Jo, I. Hypoxia and vascular endothelial growth factor acutely up-regulate angiopoietin-1 and Tie2 mRNA in bovine retinal pericytes. Microvasc. Res. 2003, 65, 125–131. [Google Scholar] [CrossRef]
  128. Yamagishi, S.; Yonekura, H.; Yamamoto, Y.; Fujimori, H.; Sakurai, S.; Tanaka, N.; Yamamoto, H. Vascular endothelial growth factor acts as a pericyte mitogen under hypoxic conditions. Lab. Investig. 1999, 79, 501–509. [Google Scholar]
  129. Engelhardt, S.; Al-Ahmad, A.J.; Gassmann, M.; Ogunshola, O.O. Hypoxia selectively disrupts brain microvascular endothelial tight junction complexes through a hypoxia-inducible factor-1 (HIF-1) dependent mechanism. J. Cell. Physiol. 2014, 229, 1096–1105. [Google Scholar] [CrossRef] [PubMed]
  130. Song, H.S.; Son, M.J.; Lee, Y.M.; Kim, W.J.; Lee, S.W.; Kim, C.W.; Kim, K.W. Oxygen tension regulates the maturation of the blood-brain barrier. Biochem. Biophys. Res. Commun. 2002, 290, 325–331. [Google Scholar] [CrossRef] [PubMed]
  131. Wang, Y.L.; Hui, Y.N.; Guo, B.; Ma, J.X. Strengthening tight junctions of retinal microvascular endothelial cells by pericytes under normoxia and hypoxia involving angiopoietin-1 signal way. Eye Lond. 2007, 21, 1501–1510. [Google Scholar] [CrossRef] [Green Version]
  132. Olsen, J.J.; Pohl, S.O.; Deshmukh, A.; Visweswaran, M.; Ward, N.C.; Arfuso, F.; Agostino, M.; Dharmarajan, A. The Role of Wnt Signalling in Angiogenesis. Clin. Biochem. Rev. 2017, 38, 131–142. [Google Scholar] [PubMed]
  133. Korovina, I.; Neuwirth, A.; Sprott, D.; Weber, S.; Sardar Pasha, S.P.B.; Gercken, B.; Breier, G.; El-Armouche, A.; Deussen, A.; Karl, M.O.; et al. Hematopoietic hypoxia-inducible factor 2alpha deficiency ameliorates pathological retinal neovascularization via modulation of endothelial cell apoptosis. FASEB J. 2019, 33, 1758–1770. [Google Scholar] [CrossRef] [PubMed]
  134. Coulon, C.; Georgiadou, M.; Roncal, C.; De Bock, K.; Langenberg, T.; Carmeliet, P. From vessel sprouting to normalization: Role of the prolyl hydroxylase domain protein/hypoxia-inducible factor oxygen-sensing machinery. Arterioscler. Thromb. Vasc. Biol. 2010, 30, 2331–2336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Xiao, Y.; Peng, H.; Hong, C.; Chen, Z.; Deng, X.; Wang, A.; Yang, F.; Yang, L.; Chen, C.; Qin, X. PDGF Promotes the Warburg Effect in Pulmonary Arterial Smooth Muscle Cells via Activation of the PI3K/AKT/mTOR/HIF-1alpha Signaling Pathway. Cell. Physiol. Biochem. 2017, 42, 1603–1613. [Google Scholar] [CrossRef] [PubMed]
  136. Suri, C.; Jones, P.F.; Patan, S.; Bartunkova, S.; Maisonpierre, P.C.; Davis, S.; Sato, T.N.; Yancopoulos, G.D. Requisite role of angiopoietin-1, a ligand for the TIE2 receptor, during embryonic angiogenesis. Cell 1996, 87, 1171–1180. [Google Scholar] [CrossRef] [Green Version]
  137. Saharinen, P.; Eklund, L.; Miettinen, J.; Wirkkala, R.; Anisimov, A.; Winderlich, M.; Nottebaum, A.; Vestweber, D.; Deutsch, U.; Koh, G.Y.; et al. Angiopoietins assemble distinct Tie2 signalling complexes in endothelial cell-cell and cell-matrix contacts. Nat. Cell Biol. 2008, 10, 527–537. [Google Scholar] [CrossRef]
  138. Watts, E.R.; Walmsley, S.R. Inflammation and Hypoxia: HIF and PHD Isoform Selectivity. Trends Mol. Med. 2019, 25, 33–46. [Google Scholar] [CrossRef] [Green Version]
  139. Holger, K.; Eltzschig, M.D.; Peter Carmeliet, M.D. Hypoxia and Inflammation. Mech. Dis. 2011, 364, 656–665. [Google Scholar]
  140. Imtiyaz, H.Z.; Simon, M.C. Hypoxia-inducible factors as essential regulators of inflammation. Curr. Top. Microbiol. Immunol. 2010, 345, 105–120. [Google Scholar] [CrossRef] [Green Version]
  141. Sormendi, S.; Deygas, M.; Sinha, A.; Krüger, A.; Kourtzelis, I.; Le Lay, G.; Bernard, M.; Sáez, P.J.; Gerlach, M.; Franke, K.; et al. HIF2α is a Direct Regulator of Neutrophil Motility. Blood 2021, 137, 3416–3427, accepted for publication. [Google Scholar] [CrossRef]
  142. Coffelt, S.B.; Wellenstein, M.D.; de Visser, K.E. Neutrophils in cancer: Neutral no more. Nat. Rev. Cancer 2016, 16, 431–446. [Google Scholar] [CrossRef] [Green Version]
  143. Liang, W.; Ferrara, N. The Complex Role of Neutrophils in Tumor Angiogenesis and Metastasis. Cancer Immunol. Res. 2016, 4, 83–91. [Google Scholar] [CrossRef] [Green Version]
  144. Kessler, D.A.; Langer, R.S.; Pless, N.A.; Folkman, J. Mast cells and tumor angiogenesis. Int. J. Cancer 1976, 18, 703–709. [Google Scholar] [CrossRef]
  145. Davis, B.P.; Rothenberg, M.E. Eosinophils and cancer. Cancer Immunol. Res. 2014, 2, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Yang, C.; Lee, H.; Pal, S.; Jove, V.; Deng, J.; Zhang, W.; Hoon, D.S.; Wakabayashi, M.; Forman, S.; Yu, H. B cells promote tumor progression via STAT3 regulated-angiogenesis. PLoS ONE 2013, 8, e64159. [Google Scholar] [CrossRef] [Green Version]
  147. De Palma, M.; Biziato, D.; Petrova, T.V. Microenvironmental regulation of tumour angiogenesis. Nat. Rev. Cancer 2017, 17, 457–474. [Google Scholar] [CrossRef] [PubMed]
  148. Kaidi, A.; Qualtrough, D.; Williams, A.C.; Paraskeva, C. Direct transcriptional up-regulation of cyclooxygenase-2 by hypoxia-inducible factor (HIF)-1 promotes colorectal tumor cell survival and enhances HIF-1 transcriptional activity during hypoxia. Cancer Res. 2006, 66, 6683–6691. [Google Scholar] [CrossRef] [Green Version]
  149. Ceradini, D.J.; Kulkarni, A.R.; Callaghan, M.J.; Tepper, O.M.; Bastidas, N.; Kleinman, M.E.; Capla, J.M.; Galiano, R.D.; Levine, J.P.; Gurtner, G.C. Progenitor cell trafficking is regulated by hypoxic gradients through HIF-1 induction of SDF-1. Nat. Med. 2004, 10, 858–864. [Google Scholar] [CrossRef] [PubMed]
  150. Zhang, W.; Petrovic, J.M.; Callaghan, D.; Jones, A.; Cui, H.; Howlett, C.; Stanimirovic, D. Evidence that hypoxia-inducible factor-1 (HIF-1) mediates transcriptional activation of interleukin-1beta (IL-1b) in astrocyte cultures. J. Neuroimmunol. 2006, 174, 63–73. [Google Scholar] [CrossRef]
  151. Oliver, K.M.; Taylor, C.T.; Cummins, E.P. Hypoxia. Regulation of NFkappaB signalling during inflammation: The role of hydroxylases. Arthritis Res. Ther. 2009, 11, 215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Mateo, J.; Garcia-Lecea, M.; Cadenas, S.; Hernandez, C.; Moncada, S. Regulation of hypoxia-inducible factor-1alpha by nitric oxide through mitochondria-dependent and -independent pathways. Biochem. J. 2003, 376, 537–544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Szade, A.; Grochot-Przeczek, A.; Florczyk, U.; Jozkowicz, A.; Dulak, J. Cellular and molecular mechanisms of inflammation-induced angiogenesis. IUBMB Life 2015, 67, 145–159. [Google Scholar] [CrossRef] [PubMed]
  154. Dulak, J.; Jozkowicz, A.; Dembinska-Kiec, A.; Guevara, I.; Zdzienicka, A.; Zmudzinska-Grochot, D.; Florek, I.; Wojtowicz, A.; Szuba, A.; Cooke, J.P. Nitric oxide induces the synthesis of vascular endothelial growth factor by rat vascular smooth muscle cells. Arterioscler. Thromb. Vasc. Biol. 2000, 20, 659–666. [Google Scholar] [CrossRef] [Green Version]
  155. Jozkowicz, A.; Cooke, J.P.; Guevara, I.; Huk, I.; Funovics, P.; Pachinger, O.; Weidinger, F.; Dulak, J. Genetic augmentation of nitric oxide synthase increases the vascular generation of VEGF. Cardiovasc. Res. 2001, 51, 773–783. [Google Scholar] [CrossRef] [Green Version]
  156. Branco-Price, C.; Zhang, N.; Schnelle, M.; Evans, C.; Katschinski, D.M.; Liao, D.; Ellies, L.; Johnson, R.S. Endothelial cell HIF-1alpha and HIF-2alpha differentially regulate metastatic success. Cancer Cell 2012, 21, 52–65. [Google Scholar] [CrossRef] [Green Version]
  157. Wu, J.B.; Tang, Y.L.; Liang, X.H. Targeting VEGF pathway to normalize the vasculature: An emerging insight in cancer therapy. Onco. Targets Ther. 2018, 11, 6901–6909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Mazzone, M.; Dettori, D.; de Oliveira, R.L.; Loges, S.; Schmidt, T.; Jonckx, B.; Tian, Y.M.; Lanahan, A.A.; Pollard, P.; de Almodovar, C.R.; et al. Heterozygous deficiency of PHD2 restores tumor oxygenation and inhibits metastasis via endothelial normalization. Cell 2009, 136, 839–851. [Google Scholar] [CrossRef] [Green Version]
  159. Fallah, J.; Rini, B.I. HIF Inhibitors: Status of Current Clinical Development. Curr. Oncol. Rep. 2019, 21, 6. [Google Scholar] [CrossRef]
  160. Lee, K.; Zhang, H.; Qian, D.Z.; Rey, S.; Liu, J.O.; Semenza, G.L. Acriflavine inhibits HIF-1 dimerization, tumor growth, and vascularization. Proc. Natl. Acad. Sci. USA 2009, 106, 17910–17915. [Google Scholar] [CrossRef] [Green Version]
  161. Mangraviti, A.; Raghavan, T.; Volpin, F.; Skuli, N.; Gullotti, D.; Zhou, J.; Asnaghi, L.; Sankey, E.; Liu, A.; Wang, Y.; et al. HIF-1alpha- Targeting Acriflavine Provides Long Term Survival and Radiological Tumor Response in Brain Cancer Therapy. Sci. Rep. 2017, 7, 14978. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Kong, D.; Park, E.J.; Stephen, A.G.; Calvani, M.; Cardellina, J.H.; Monks, A.; Fisher, R.J.; Shoemaker, R.H.; Melillo, G. Echinomycin, a small-molecule inhibitor of hypoxia-inducible factor-1 DNA-binding activity. Cancer Res. 2005, 65, 9047–9055. [Google Scholar] [CrossRef] [Green Version]
  163. Thomas, A.; Samykutty, A.; Gomez-Gutierrez, J.G.; Yin, W.; Egger, M.E.; McNally, M.; Chuong, P.; MacCuaig, W.M.; Albeituni, S.; Zeiderman, M.; et al. Actively Targeted Nanodelivery of Echinomycin Induces Autophagy-Mediated Death in Chemoresistant Pancreatic Cancer In Vivo. Cancers 2020, 12, 2279. [Google Scholar] [CrossRef] [PubMed]
  164. Hur, E.; Kim, H.H.; Choi, S.M.; Kim, J.H.; Yim, S.; Kwon, H.J.; Choi, Y.; Kim, D.K.; Lee, M.O.; Park, H. Reduction of hypoxia-induced transcription through the repression of hypoxia-inducible factor-1alpha/aryl hydrocarbon receptor nuclear translocator DNA binding by the 90-kDa heat-shock protein inhibitor radicicol. Mol. Pharmacol. 2002, 62, 975–982. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Ban, H.S.; Kim, B.K.; Lee, H.; Kim, H.M.; Harmalkar, D.; Nam, M.; Park, S.K.; Lee, K.; Park, J.T.; Kim, I.; et al. The novel hypoxia-inducible factor-1alpha inhibitor IDF-11774 regulates cancer metabolism, thereby suppressing tumor growth. Cell Death Dis. 2017, 8, e2843. [Google Scholar] [CrossRef]
  166. Zhang, P.C.; Liu, X.; Li, M.M.; Ma, Y.Y.; Sun, H.T.; Tian, X.Y.; Wang, Y.; Liu, M.; Fu, L.S.; Wang, Y.F.; et al. AT-533, a novel Hsp90 inhibitor, inhibits breast cancer growth and HIF-1alpha/VEGF/VEGFR-2-mediated angiogenesis in vitro and in vivo. Biochem. Pharmacol. 2020, 172, 113771. [Google Scholar] [CrossRef]
  167. Mabjeesh, N.J.; Escuin, D.; LaVallee, T.M.; Pribluda, V.S.; Swartz, G.M.; Johnson, M.S.; Willard, M.T.; Zhong, H.; Simons, J.W.; Giannakakou, P. 2ME2 inhibits tumor growth and angiogenesis by disrupting microtubules and dysregulating HIF. Cancer Cell 2003, 3, 363–375. [Google Scholar] [CrossRef] [Green Version]
  168. Kulke, M.H.; Chan, J.A.; Meyerhardt, J.A.; Zhu, A.X.; Abrams, T.A.; Blaszkowsky, L.S.; Regan, E.; Sidor, C.; Fuchs, C.S. A prospective phase II study of 2-methoxyestradiol administered in combination with bevacizumab in patients with metastatic carcinoid tumors. Cancer Chemother. Pharmacol. 2011, 68, 293–300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Zhou, S.; Wang, F.; Wong, E.T.; Fonkem, E.; Hsieh, T.C.; Wu, J.M.; Wu, E. Salinomycin: A novel anti-cancer agent with known anti-coccidial activities. Curr. Med. Chem. 2013, 20, 4095–4101. [Google Scholar] [CrossRef] [Green Version]
  170. Dewangan, J.; Srivastava, S.; Mishra, S.; Divakar, A.; Kumar, S.; Rath, S.K. Salinomycin inhibits breast cancer progression via targeting HIF-1alpha/VEGF mediated tumor angiogenesis in vitro and in vivo. Biochem. Pharmacol. 2019, 164, 326–335. [Google Scholar] [CrossRef]
  171. Okuno, T.; Kawai, K.; Hata, K.; Murono, K.; Emoto, S.; Kaneko, M.; Sasaki, K.; Nishikawa, T.; Tanaka, T.; Nozawa, H. SN-38 Acts as a Radiosensitizer for Colorectal Cancer by Inhibiting the Radiation-induced Up-regulation of HIF-1alpha. Anticancer Res. 2018, 38, 3323–3331. [Google Scholar] [CrossRef] [PubMed]
  172. Choueiri, T.K.; Kaelin, W.G., Jr. Targeting the HIF2-VEGF axis in renal cell carcinoma. Nat. Med. 2020, 26, 1519–1530. [Google Scholar] [CrossRef] [PubMed]
  173. Gaete, D.; Rodriguez, D.; Watts, D.; Sormendi, S.; Chavakis, T.; Wielockx, B. HIF-Prolyl Hydroxylase Domain Proteins (PHDs) in Cancer-Potential Targets for Anti-Tumor Therapy? Cancers 2021, 13, 988. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Hypoxia pathway proteins–a regulatory network. Under normoxic conditions, PHDs 1–3 can preferentially hydroxylate HIF1α or 2α, which leads to their VHL-mediated degradation. FIH preferentially hydroxylates HIF1α rather than HIF2α, preventing necessary interaction with co-factors [6,8]. Additionally, LIMD1, a scaffold protein that helps assemble the complex with VHL [9,10] and numerous other regulators of PHD expression/activity have been described (green box—Normoxia [11,12,13,14,15,16,17]). During hypoxia-associated events, including inflammation or tumor growth, HIF1 is activated (green box—Hypoxia), on the other hand, HIF1 supports the transcription of PHD, LIMD1, and Nitric Oxide synthase (NOS), which in turn inhibit HIF1 activity. HIF2 can induce the expression of Arginase 1, which competes with NOS for the same substrate, thus reducing NO production efficiency and reducing HIF1 inactivation as well [9,10,18,19,20,21]. HIF1α-AS2 promotes HIF1 transcription only in the first few hours after hypoxia induction but subsequently inhibits HIF1 activity; in contrast, HIF1α-AS2 continues to promote HIF2 expression even after several hours of hypoxia [4,22].
Figure 1. Hypoxia pathway proteins–a regulatory network. Under normoxic conditions, PHDs 1–3 can preferentially hydroxylate HIF1α or 2α, which leads to their VHL-mediated degradation. FIH preferentially hydroxylates HIF1α rather than HIF2α, preventing necessary interaction with co-factors [6,8]. Additionally, LIMD1, a scaffold protein that helps assemble the complex with VHL [9,10] and numerous other regulators of PHD expression/activity have been described (green box—Normoxia [11,12,13,14,15,16,17]). During hypoxia-associated events, including inflammation or tumor growth, HIF1 is activated (green box—Hypoxia), on the other hand, HIF1 supports the transcription of PHD, LIMD1, and Nitric Oxide synthase (NOS), which in turn inhibit HIF1 activity. HIF2 can induce the expression of Arginase 1, which competes with NOS for the same substrate, thus reducing NO production efficiency and reducing HIF1 inactivation as well [9,10,18,19,20,21]. HIF1α-AS2 promotes HIF1 transcription only in the first few hours after hypoxia induction but subsequently inhibits HIF1 activity; in contrast, HIF1α-AS2 continues to promote HIF2 expression even after several hours of hypoxia [4,22].
Ijms 22 09191 g001
Figure 2. Sprouting angiogenesis is related to hypoxic signaling: (A) Cells under hypoxic conditions release angiogenic factors that help loosen the pericytes, resulting in unstable vessels and activation of VEGFR2-expressing tip cells. (B) Pericytes detach, EC becomes hyper-permeable in response to VEGF [37], metalloproteases extravasate to degrade the ECM causing release of stored growth factors and proenzymes. A new, provisional ECM is made. Endothelial cells start invading the provisional ECM [116]. (C) Hypoxia establishes Tip/Stalk cells. Tip cells are highly mobile, while stalk cells are immotile and proliferative, enabling pericyte recruitment while the branch extends towards hypoxic signals [117]. Pericyte recruitment stabilizes the vessel. Furthermore, cells in the termination area release Vasohibin1, leading to angiogenesis inhibition and vessel stabilization [112]. (D) The angiogenic vessel fuses with another vessel. Oxygenated blood circulates through the lumen, dampening the hypoxic signals. Mesenchymal cells migrate along the vessel and form mature pericytes, stabilizing the vessel and helping it mature [109].
Figure 2. Sprouting angiogenesis is related to hypoxic signaling: (A) Cells under hypoxic conditions release angiogenic factors that help loosen the pericytes, resulting in unstable vessels and activation of VEGFR2-expressing tip cells. (B) Pericytes detach, EC becomes hyper-permeable in response to VEGF [37], metalloproteases extravasate to degrade the ECM causing release of stored growth factors and proenzymes. A new, provisional ECM is made. Endothelial cells start invading the provisional ECM [116]. (C) Hypoxia establishes Tip/Stalk cells. Tip cells are highly mobile, while stalk cells are immotile and proliferative, enabling pericyte recruitment while the branch extends towards hypoxic signals [117]. Pericyte recruitment stabilizes the vessel. Furthermore, cells in the termination area release Vasohibin1, leading to angiogenesis inhibition and vessel stabilization [112]. (D) The angiogenic vessel fuses with another vessel. Oxygenated blood circulates through the lumen, dampening the hypoxic signals. Mesenchymal cells migrate along the vessel and form mature pericytes, stabilizing the vessel and helping it mature [109].
Ijms 22 09191 g002
Table 1. Overview of hypoxia-related proteins involved in angiogenesis.
Table 1. Overview of hypoxia-related proteins involved in angiogenesis.
ProteinFunctionHIF PathwayType of Regulation by HypoxiaReference
ANG1, ANG2Cell-cell adhesionHIF1 for ANG1;
HIF1 and HIF2 for ANG2
Up or down[2,68,69,70,71]
Wnt/FzdAngiogenesis initiation Pericyte recruitment Inflammation regulationHIF1 and HIF2; independent mechanismsUp or down[45,72,73,74,75,76,77]
VEGF familyAngiogenesis initiation and patterning.
Tip cell formation
Vasculature patterning
HIF1 and HIF2Up[78,79,80]
SemaphorinsVasculature patterningHIF1 for Sema4DUp[81,82,83,84]
NeuropilinsCo-receptor for VEGF;
binds semaphorins;
Vasculature patterning
HIF1 and HIF2Up or down depending on cell location and type[81,85,86,87,88]
Ephrins/EphVascular patterningHIF1Up[89,90]
Netrin1/UNC5 and DCC.Vascular patterningHIF1Up[91,92]
Slits/ROBOVascular patterning Up[93]
Dll4/NOTCHStalk/tip cell specification;
decision to sprout or widen a vessel;
arterial specification;
EC quiescence and survival
HIF2 for Dll4
FIH for Notch
Up[1,65,94,95,96,97]
Transforming growth factor β (TGFβ)Regulation of angiogenic factors and EC growth;
vessel maturation (including ECM formation and smooth muscle cell recruitment);
angiogenesis resolution.
TGFβ1 can inhibit PHD2Up in most cases but sometimes down[98,99,100,101]
Bone Morphogenic proteins (BMP)Increases angiogenesis;
apoptosis regulation;
Arterial versus venous specification
Indirect:
hypoxia inhibits BMP-binding endothelial regulator, which inhibits BMP expression
Up[102]
TGFβ receptors type I and IIEach of the different members of the TGFβ family binds a different combination of TGFβ type I and II receptorsIndirect: HIF1 mediates increase of EZH2, which inhibits TGFβRII expressionUp of Alk1 but not Alk5
Down of TGFβRII in prostate cancer.
[102,103,104]
TGFβ receptors type III
(Betaglycan and Endoglin)
These are accessory receptors for the TGFβ family. They have a more indirect role in TGFβ signal induction than type I or II receptorsHIF1 for EndoglinUp for Endoglin
Down for betaglycan in rat lungs
[105,106]
Cited2Vascularization of placentaCited2 can inhibit HIF-1α signalingUp[107,108]
Matrix metalloproteinasesDegradation of the ECMHIF1 and HIF2Up[54,55]
PDGFB/PDGFBRRecruitment of mural cellsHIF1Up[2,109,110,111]
Vasohibin1Angiogenesis termination
Angiogenesis inhibition
Indirect: Hypoxia induces VEGF release, which induces vasohibin expression
Vasohibin1 can induce PHD-mediated degradation of HIF1
Up[112,113,114]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rodriguez, D.; Watts, D.; Gaete, D.; Sormendi, S.; Wielockx, B. Hypoxia Pathway Proteins and Their Impact on the Blood Vasculature. Int. J. Mol. Sci. 2021, 22, 9191. https://doi.org/10.3390/ijms22179191

AMA Style

Rodriguez D, Watts D, Gaete D, Sormendi S, Wielockx B. Hypoxia Pathway Proteins and Their Impact on the Blood Vasculature. International Journal of Molecular Sciences. 2021; 22(17):9191. https://doi.org/10.3390/ijms22179191

Chicago/Turabian Style

Rodriguez, Diego, Deepika Watts, Diana Gaete, Sundary Sormendi, and Ben Wielockx. 2021. "Hypoxia Pathway Proteins and Their Impact on the Blood Vasculature" International Journal of Molecular Sciences 22, no. 17: 9191. https://doi.org/10.3390/ijms22179191

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