Next Article in Journal
Members of the GADD45 Protein Family Show Distinct Propensities to form Toxic Amyloid-Like Aggregates in Physiological Conditions
Next Article in Special Issue
Dietary Intake of 17α-Ethinylestradiol Promotes HCC Progression in Humanized Male Mice Expressing Sex Hormone-Binding Globulin
Previous Article in Journal
Fucose Ameliorates Tritrichomonas sp.-Associated Illness in Antibiotic-Treated Muc2−/− Mice
Previous Article in Special Issue
Hydrogen Sulfide Reduces Ischemia and Reperfusion Injury in Neuronal Cells in a Dose- and Time-Dependent Manner
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Chronic and Cycling Hypoxia: Drivers of Cancer Chronic Inflammation through HIF-1 and NF-κB Activation: A Review of the Molecular Mechanisms

by
Jan Korbecki
1,
Donata Simińska
1,
Magdalena Gąssowska-Dobrowolska
2,
Joanna Listos
3,
Izabela Gutowska
1,
Dariusz Chlubek
1 and
Irena Baranowska-Bosiacka
1,*
1
Department of Biochemistry and Medical Chemistry, Pomeranian Medical University in Szczecin, Powstańców Wielkopolskich 72 Av., 70-111 Szczecin, Poland
2
Department of Cellular Signalling, Mossakowski Medical Research Institute, Polish Academy of Sciences, Pawińskiego 5, 02-106 Warsaw, Poland
3
Department of Pharmacology and Pharmacodynamics, Medical University of Lublin, Chodźki 4a St., 20-093 Lublin, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(19), 10701; https://doi.org/10.3390/ijms221910701
Submission received: 15 September 2021 / Revised: 28 September 2021 / Accepted: 1 October 2021 / Published: 2 October 2021

Abstract

:
Chronic (continuous, non-interrupted) hypoxia and cycling (intermittent, transient) hypoxia are two types of hypoxia occurring in malignant tumors. They are both associated with the activation of hypoxia-inducible factor-1 (HIF-1) and nuclear factor κB (NF-κB), which induce changes in gene expression. This paper discusses in detail the mechanisms of activation of these two transcription factors in chronic and cycling hypoxia and the crosstalk between both signaling pathways. In particular, it focuses on the importance of reactive oxygen species (ROS), reactive nitrogen species (RNS) together with nitric oxide synthase, acetylation of HIF-1, and the action of MAPK cascades. The paper also discusses the importance of hypoxia in the formation of chronic low-grade inflammation in cancerous tumors. Finally, we discuss the effects of cycling hypoxia on the tumor microenvironment, in particular on the expression of VEGF-A, CCL2/MCP-1, CXCL1/GRO-α, CXCL8/IL-8, and COX-2 together with PGE2. These factors induce angiogenesis and recruit various cells into the tumor niche, including neutrophils and monocytes which, in the tumor, are transformed into tumor-associated neutrophils (TAN) and tumor-associated macrophages (TAM) that participate in tumorigenesis.

Graphical Abstract

1. Introduction

The growing knowledge of tumors indicates the significance of the tumor microenvironment, a collection of factors that act on cancer cells in the tumor. These factors include tumor-associated cells [1,2] along with elements of intercellular signaling, such as growth factors [3], lipid mediators [4], chemokines [5], and many others. Nutrient levels, lactic acid concentration, and acidification of the tumor microenvironment [6], as well as reduced oxygen levels, i.e., hypoxia, associated with tumor growth, are also important for tumor growth. Hypoxic conditions significantly alter the functioning of cancer cells as well as tumor-associated cells.
An important aspect of hypoxia in the tumor microenvironment is chronic low-grade inflammation. The role of inflammation supports the fight of the immune system against pathogens and is an element strengthening the anti-tumor response [7]. However, inflammatory processes also include mechanisms that inhibit the body from unduly responding to pro-inflammatory factors. They protect tissues from damage by their own over-reactive mechanisms designed to fight pathogens. During chronic inflammation, these mechanisms lead to the inhibition of anti-tumor response [8] and thus promote cancerous tumor growth [9].
This review expands on the mechanisms of the activation of hypoxia-inducible factors (HIFs) and nuclear factor κB (NF-κB) presented in our previous reviews on the effects of hypoxia on the CC [10] and CXC [11] sub-family chemokine systems. These papers show the exact mechanisms responsible for the induction of the expression of individual chemokines by chronic and cycling hypoxia. In this paper, we focus on the activation of HIFs and NF-κB by different types of hypoxia and the crosstalk between the activation pathways of these two transcription factors.

2. Chronic Hypoxia

2.1. Activation of the Hypoxia-Inducible Factor by Oxygen Reduction: The Role of Hydroxylation

The intense division of cancer cells results in the proliferation of tumor tissue. This process does not go hand in hand with angiogenesis, i.e., the formation of new blood vessels. In this way, due to the low availability of blood vessels, the tumor has areas with chronically reduced oxygen concentration. This microenvironment is called chronic (continuous, non-interrupted) hypoxia.
The most important and best-known proteins activated in hypoxia are three hypoxia-inducible factors (HIF-1, HIF-2, and HIF-3). The first two, HIF-1 and HIF-2, are responsible for the transcription of genes induced by hypoxia, while HIF-3, in addition to inducing gene expression, also decreases the activity of HIF-1 and HIF-2 [12,13,14].
All three HIFs are composed of two subunits, alpha and beta. The HIF-β subunits, also known as aryl hydrocarbon nuclear translocators (ARNT), are not regulated by any changes in oxygen, although a study on high-risk multiple myeloma cells shows that chronic hypoxia increases HIF-1β expression via NF-κB [15]. The highest expression of HIF-2β occurs in the brain and kidneys [16]. HIF-2β interferes with the function of HIF-1 and is, therefore, a suppressor gene in cancers such as oral squamous cell carcinoma [17], non-small cell lung cancer [18] and hepatocellular carcinoma [19].
In contrast to HIF-β subunits, the expression levels of HIF-1α, HIF-2α, and HIF-3α subunits are tightly regulated by changes in oxygen concentration through proteolytic degradation and transcriptional regulation. In addition, HIF-3α expression is upregulated by HIF-1 and HIF-2 [14]. This represents one of the many mechanisms of self-regulation of HIF transcriptional activity.
In normoxia, HIF-α undergoes hydroxylation on the proline residue in the N-terminal oxygen-dependent degradation domain (NODD) and C-terminal oxygen-dependent degradation domain (CODD) by three isoforms of prolyl hydroxylase (PHD) [20,21]—oxygen-dependent enzymes with an iron atom in the catalytic center [22]. PHD2 and PHD3 have similar rates of catalysis, while PHD1 has a three times lower rate than the remaining two PHDs [23].
PHDs induce the hydroxylation of the proline residues Pro402 HIF-1α, Pro564 HIF-1α, Pro405 HIF-2α, and Pro531 HIF-2α [24]. This leads to the ubiquitination of the hydroxylated HIF-α subunits by the von Hippel–Lindau protein (pVHL) [22,25,26,27], followed by the proteasomal degradation of HIF-1α and HIF-2α by 26S proteasome [28,29]. In the absence of HIF-1α and HIF-2α in the cell, active transcriptional complexes with HIF-1β and HIF-2β are not assembled.
Another factor involved in the regulation of HIF-α transcriptional activity is the factor inhibiting HIF (FIH), an oxygen-dependent enzyme with asparaginyl hydroxylase activity for HIF-α subunits. This enzyme causes hydroxylation of HIF-α at the Asn803 HIF-1α and Asn847 HIF-2α residues [30,31]. This hydroxylation inhibits the interaction of the HIF-α subunit with CBP/p300 [30,32,33]. The interaction of HIF-α with this coactivator is necessary for the transcription of HIF-dependent genes. Therefore, FIH action provides a mechanism for reducing the transcriptional activity of HIFs in normoxia.
FIH and PHD require different levels of oxygen to fully function. The activity of PHD is significantly reduced when the oxygen concentration in the cell’s environment is reduced to 5% [23]. The Michaelis constant (Km) for these enzymes relative to substrate oxygen is 230–250 μM [34]. In hypoxia, there is an increase in PHD expression by HIFs which increases the activity of these enzymes during reoxygenation [35,36,37,38]. In contrast, FIH requires half the oxygen concentration necessary for PHD activity [39]. In this way, FIHs inhibit the transcriptional activity of HIFs at oxygen concentrations where PHD activity is already reduced.
The reduction in PHD’s activity results in a decrease in the level of hydroxylation of the proline residue on HIF-α. This leads to (1) a decrease in the degradation of HIF-α, (2) the accumulation of these proteins in the cell, (3) the dimerization of the corresponding HIF-α and HIF-β subunits, and finally, (4) production of HIF-1, HIF-2, and HIF-3 transported to the cell nucleus. The hydroxylation of HIF-α by FIH does not occur at low oxygen concentrations. This leads to an interaction of the HIF-α subunit with CBP/p300 on the promoters of genes with hypoxia response element (HRE) sequences [30,32,33] and then to the increased expression of hypoxia-dependent genes.
The accumulation of individual HIF-α—and so, the activation of individual HIFs—depends on the duration of hypoxia [40]. HIF-1 is activated in the first 4 h of chronic hypoxia, after which HIF-1α protein levels decrease [40,41]. In contrast, maximum HIF-2α and HIF-3α levels occur after 24–48 h of hypoxia [40]. This is associated with an increased expression of hypoxia-associated factor (HAF), which causes pVHL-independent proteolytic degradation of HIF-1α [41]. In prolonged chronic hypoxia, reduced HIF-1α expression may also be caused by the activity of heat shock protein 70 (Hsp70), which, together with the carboxyl terminus of Hsc70-interacting protein (CHIP), causes the ubiquitination of HIF-1α but not HIF-2α [42].
It should be noted that the expression of HIF-2α varies in different tumors. It is absent in small cell lung carcinoma, while it is present in non-small cell lung carcinoma [43]. Additionally, an in vivo study shows that in tumor cells, the levels of HIF-1α and HIF-2α are high on average but vary depending on the type of cells [44]. In tumor-associated macrophages (TAM), HIF-2α [44,45] and HIF-1α [46] levels are high.

2.2. Acetylation of HIF-α as a Possible Mechanism for the Regulation of HIF’s Activity in Chronic Hypoxia

Hydroxylation is not the only mechanism that can alter HIF-α stability. Another post-translational modification that regulates the stability of HIF-α is acetylation. HIF-1α has 12 amino acid residues that are potentially subject to ubiquitination [47]. Depending on which of these residues is acetylated, this post-translational modification may either increase or decrease the stability and transcriptional activity of this HIF-1 subunit. This process has been thoroughly described for the chronic hypoxia model. Nevertheless, the effect of acetylation on HIF-1 activity in the model of cycling hypoxia is poorly understood.
In chronic hypoxia, protein 14-3-3ζ promotes the interaction of histone deacetylase (HDAC)4 with HIF-1α, which reduces the acetylation of this HIF-1 subunit [48]. As a consequence, the stability of the HIF-1α protein is increased. This mechanism has been demonstrated in a hepatocellular carcinoma model [48]. Increased stability and transcriptional activities of HIF-1α have also been observed in HDAC1, HDAC3 [49], HDAC4 [50,51,52], HDAC5 [51], and HDAC6 [50]. Importantly, HDAC4 and HDAC5 bind to HIF-1α, which prevents the hydroxylation of this subunit of HIF-1 via FIH [51]. In chronic hypoxia, HDAC7 forms a complex with HIF-1α in the cell nucleus, which increases the transcriptional activity of HIF-1 [53]. In contrast, acetylation of Lys532 HIF-1α by arrest defective 1 (ARD1) reduces the stability and transcriptional activity of HIF-1α [54]. In hypoxia, ARD1 expression is decreased, which increases HIF-1 activation.
Acetylation of HIF-1α may also increase the stability and transcriptional activity of this HIF subunit in chronic hypoxia. HIF-1α, but not HIF-2α, undergoes acetylation at the Lys709 residue by p300, which increases the stability of HIF-1α [55]. HIF-1α has 12 residues that undergo ubiquitination [47]. One of these is the Lys709 residue. Acetylation of this residue prevents its ubiquitination; this results in an increase in the stability of HIF-1α. In addition, the deacetylation of the Lys674 residue of HIF-1α in normoxia by sirtuin (SIRT)1 blocks recruitment of this HIF-1 subunit from p300 [56]. In hypoxia, a decrease in SIRT1 activity causes acetylation of Lys674 HIF-1α by p300/CBP-associated factor (PCAF). Other sirtuins also reduce HIF-1 pathway activation, including SIRT2 [57], SIRT3 [58], and SIRT7 [59].

2.3. The Role of ROS and NO in the Activation of HIFs during Chronic Hypoxia

An important part of the cellular response to hypoxia are reactive oxygen species (ROS), which increase HIF-1 stability. In chronic hypoxia, this process is much less important than the effects of ROS on signaling pathways in cycling hypoxia. Chronic hypoxia is associated with an increase in ROS generation by complex III of the mitochondrial electron transport chain [60,61,62]. Circadian locomotor output cycle protein kaput (CLOCK) may also be responsible for increasing ROS levels in chronic hypoxia [63]. ROS increase the activation of HIF-1 and NF-κB. Specifically, in the cytoplasm, ROS inhibit the activity of PHD [62,64] and FIH [31]. Importantly, the changes that ROS cause in these enzymes vary. FIH is more sensitive to ROS but is inactivated more permanently than PHD [31]. In the case of PHD, it has been suggested that ROS cause oxidation of the iron atom, important in the activity of these enzymes that regulate HIF-α stability and function [65]; no mechanism has been established for FIH [31]. At the same time, ROS activates NF-κB through various mechanisms [63,66,67], and subsequently, NF-κB increases the expression of HIF-1α mRNA. In chronic hypoxia, there is also an increase in ROS generation in cytoplasm. HIF-1 can increase NADPH oxidase (NOX)4 expression and thus ROS generation in the cytoplasm [68,69], although it is possible that, in hypoxia, NOX4 expression is increased directly by NF-κB p65/RelA [70]. Significantly, the relevance of mitochondrial ROS for HIF-1α stability is disputed by some researchers [71]. The inhibition of PHD by ROS during chronic hypoxia is also questioned [71].
Activation of HIFs is dependent on nitric oxide (NO) levels (Figure 1). This is important because many cancers are accompanied by an increase in inducible nitric oxide synthase (iNOS) expression and an increase in NO production, resulting in a poorer prognosis for the patient [72,73,74]. HIF-1α is S-nitrosylated on Cys533 by NO, resulting in increased stability of this protein in normoxia [75]. NO also binds to the iron atom at the catalytic center in PHD, which inhibits the activity of this enzyme [76]. On the other hand, in hypoxia, NO can interfere with HIF-1 activation and function [77]. In combination with ROS, NO increases the concentration of calcium ions in the cytoplasm which activates calpain [78]—a protease that degrades HIF-1α, independently of the 26S proteasome. In chronic hypoxia, NO also restores the activity of enzymes involved in HIF-α hydroxylation [64,79] in a mechanism dependent on the interaction between NO with ROS.

2.4. MAPK and AP-1 Kinases in Chronic Hypoxia

During hypoxia, mitogen-activated protein kinase (MAPK) cascades are activated and play an important role in the cellular response to reduced oxygen concentration. These processes have been thoroughly researched in the chronic hypoxia model as opposed to cycling hypoxia. Nevertheless, the activation of MAPK cascades is similar in both hypoxia, and therefore, in all likelihood, the molecular mechanisms described in this section reflect HIF-1 activation in cycling hypoxia.
The activation of MAPK cascades in chronic hypoxia is dependent on an increased concentration of calcium ions or ROS [80,81]. An influx of calcium ions in hypoxia is caused by the opening of the L-type voltage gated Ca2+ channels [82]. ROS or calcium ions activate the p38 MAPK [61,81,82,83,84], extracellular signal-regulated kinase (ERK) MAPK [81,82,85] and c-Jun N-terminal kinase (JNK) MAPK [81,86]. This results in the activation of c-Jun and JunB, but also a decrease in the levels of c-Fos, and JunD [87]. However, it is not only that the activation of MAPK cascades activates HIFs; in a reverse process, HIF-1 may influence the activation of MAPK cascades. The increase in c-Jun expression in chronic hypoxia in mouse embryonic fibroblasts is HIF-1 dependent [88], similar to the activation of the ERK MAPK cascade [89]. c-Jun and JunB are elements of activating protein-1 (AP-1), a transcription factor that plays an important role in the expression of various genes in hypoxia.
MAPK cascades can also affect the activation and transcriptional activity of HIF-1 (Figure 2). Chronic hypoxia induces the activation of ERK MAPK [85], p38 MAPK [90], and JNK MAPK [91]. ERK MAPK causes phosphorylation on Ser641 and Ser643 of HIF-1α. This post-translational modification of HIF-1α is important for the accumulation of this subunit in the cell nucleus and interaction of this subunit with the p300 coactivator [92,93]. Additionally, HIF-1α is phosphorylated by p38 MAPK, which increases the stability of this HIF-1 subunit [90]. Another mechanism affecting HIF-1 activation is the phosphorylation of the seven in absentia homolog 2 (SIAH2) at the Thr24 and Ser29 residues by p38 MAPK [94]. This is followed by the degradation of PHD3, an enzyme responsible for hydroxylation and degradation of HIF-1α. Activation of p38 MAPK results in decreased degradation of HIF-1α.
Chronic hypoxia increases the expression of dual specificity protein phosphatase-(DUSP)-1/mitogen-activated protein kinase phosphatase 1 (MKP1) [95]. The increased expression of DUSP1 in neurons in chronic hypoxia is dependent on neuronal nitric oxide synthase (nNOS) and NO produced by nNOS [96]. NO inactivates DUSP1; in contrast, protein kinase C (PKC)ζ is responsible for increasing DUSP1 expression in fibroblasts under chronic hypoxia [97]. However, DUSP1 expression in chronic hypoxia may also be dependent on p38 MAPK, as shown by experiments on pheochromocytoma cells [98]. The use of cobalt chloride or deferoxamine (both these compounds being PHD inhibitors) showed that HIF activation may result in increased DUSP1 expression [98], although, to date, there has been no proof of a direct effect of HIF on the expression of DUSP1. Cobalt chloride and deferoxamine may also inhibit the activity of histone demethylase which leads to an altered expression of many genes [99,100]. In chronic hypoxia, DUSP2 expression is also decreased by HIF-1 [101,102]. DUSP1 and DUSP2 are enzymes that inactivate MAPK kinases (ERK MAPK and p38 MAPK) by their dephosphorylation [103]. The upregulation of DUSP1/MKP1 expression is a mechanism that protects against excessive HIF-1 activation by ERK MAPKs in chronic hypoxia.
MAPK cascades enhance the stability and transcriptional activity of HIF-1α in hypoxia. Similar mechanisms also occur in tumor cells in normoxia, where the activation of MAPK cascades also occurs [104], in particular as a result of exposure to various growth factors [105,106].
MAPK kinases in hypoxia also cause NF-κB activation. However, which MAPK cascade is responsible for this varies from model to model. In macrophages, NF-κB activation is induced by ERK cascade [107], while in Hey-A8 human ovarian carcinoma cells, it is by p38 MAPK cascade [83]. In addition, ERK MAPK causes phosphorylation of Ser276 p65/RelA NF-κB which results in the activation of this transcription factor [89].

2.5. NF-κB Activation during Chronic Hypoxia Is Important for the Full Activation of HIFs

In hypoxia, the transcriptional activity of HIFs is increased by decreasing hydroxylation and degradation of HIF-α. However, maximal HIF activation requires the activation of other pathways. In particular, NF-κB is activated during the first hours of chronic hypoxia [40]. This leads to an increase in HIF-1α mRNA expression due to the presence of an NF-κB binding site in the HIF1A gene promoter [66,108,109,110,111]. Therefore, the activation of p50 NF-κB and p65/RelA NF-κB, but not c-Rel NF-κB, results in increased HIF-1α mRNA expression [112,113,114]. HIF-1β expression is also directly upregulated by NF-κB in chronic hypoxia [15,115]. The relationship of these two transcription factors is important in hypoxia because NF-κB is activated by low oxygen concentration via multiple mechanisms [116]. Like HIF-1α, the IκB kinase β subunit (IKKβ) is also hydroxylated on Pro191 by PHD1 [117,118]. This leads to ubiquitination of the Lys63 residue of IKKβ by pVHL [119]. This post-translational modification prevents transforming growth factor (TGF)-β-activated kinase 1 (TAK1) from attaching to IKKβ. This decreases IKKβ activity and thus reduces NF-κB activation. However, ubiquitination of IKKβ does not lead to proteolytic degradation of this protein [119]. In chronic hypoxia, there is a decrease in PHD1 activity, which results in a decrease in IKKβ hydroxylation by this enzyme and, consequently, in an increase in IKKβ activity. PHD2 also plays an important role in regulating the NF-κB activation pathway [120]. This enzyme indirectly regulates phosphorylation of the inhibitor of NF-κB α subunit (IκBα). Additionally, in chronic hypoxia, there is an increase in calcium ion levels in cytoplasm, which results in activation of calcium/calmodulin-dependent kinase 2 (CaMK2). This leads to ubiquitination of Lys63 Nemo/IKKγ by ubiquitin-conjugating enzyme 13 (Ubc13) [121]. This results in the activation of IKK.
Chronic hypoxia also affects other components of the NF-κB activation pathway. FIH can catalyze the hydroxylation of IκBα (Figure 3) [122]. Nevertheless, this has no effect and is not relevant for NF-κB activation in chronic hypoxia. Activation of CaMK2 causes Lys21 IκBα sumoylation by Sumo-2/3 and consequently prevents IκBα ubiquitination [121]. This leads to the release and activation of NF-κB in the absence of IκBα degradation [121]. The MAPK cascades are also activated in hypoxia, resulting in NF-κB activation. Depending on the model, ERK MAPK, which phosphorylates Ser276 p65/RelA NF-κB, is responsible for NF-κB activation in macrophages and primary mouse keratinocytes [89,107]. In contrast, in ovarian carcinoma cell line Hey-A8 [83] and lung adenocarcinoma A549 cells [123], the p38 MAPK cascade is responsible for NF-κB activation. In addition to these mechanisms, NF-κB activation occurs in hypoxia via the phosphatidylinositol 3-kinase (PI3K)→Akt/protein kinase B (PKB) pathway [83]. This pathway is activated either by ROS or by activation of membrane receptors [83]. Additionally, this pathway can be activated in hypoxia by 14-3-3ζ, as shown in hepatocellular carcinoma [124]. Akt/PKB kinase phosphorylates p65/RelA NF-κB cause activation of this transcription factor, but this process is independent of IκBα degradation [83]. Akt/PKB can also activate mTOR, which phosphorylates Thr23 IKKα and Thr559 and Ser634 IKKβ [125]. This induces IKK activation, which leads to IκBα phosphorylation and degradation. Another pathway of NF-κB activation in chronic hypoxia is the activation of protein kinase D2 (PRKD2) [126], although the exact mechanism of activation of this kinase has not been thoroughly researched.
After activation, NF-κB forms a complex with its coactivators. These complexes include PHD2 [127] and PHD3 [128]. The property of these PHDs does not depend on enzymatic activity. This is important in inflammatory reactions and in a cell’s response to chronic hypoxia. In hypoxia, PHD expression is increased in a HIF-dependent manner [35,36,37,38], which may increase the transcriptional activity of NF-κB and hence increase HIF-1α expression.
HIF-1 and HIF-2 also increase the expression of p65/RelA NF-κB in macrophages [129]. HIF-1 can also activate NF-κB indirectly by increasing the expression of alarmin receptors [130] activated by damage-associated molecular patterns (DAMPs), i.e., molecules secreted from cells during necrosis, e.g., in an environment with hypoxia. Activation of alarmin receptors results in the activation of NF-κB. There is also an increased expression of thioredoxin reductase 1 (TrxR1) [131]. This increases the level of ROS in the cytoplasm, which leads to increased activation of NF-κB [132,133].

2.6. Inhibition of Inflammatory Responses by Chronic Hypoxia

Hypoxia is associated with a decrease in PHD1 activity, which leads to a decrease in the hydroxylation of IKKβ [117,118] and HIF-α [20,21]. This activates this kinase which leads to NF-κB activation and an increase in HIF-1α mRNA expression by NF-κB. This is followed by transcription of hypoxia-induced genes but also some pro-inflammatory genes [40]. However, also in inflammatory responses, including the action of lipopolysaccharide (LPS), PHD is inactivated, and thereby, NF-κB is activated, which induces expression of pro-inflammatory genes and increases HIF-1α expression [134,135]. However, both pathways, i.e., NF-κB and HIF, if activated simultaneously, will be mutually exclusive (Figure 4) [136,137,138]. This is because the transcription of genes induced by hypoxia (HIF) and inflammation (NF-κB) requires the coactivator p300 [139,140] and both transcription factors compete with each other for this coactivator. Additionally, there are mechanisms of inhibition of the NF-κB pathway by the HIF activation pathway, which is important in reducing overly intense inflammatory responses as well as reducing inflammatory responses in chronic hypoxia [137,138]. TAK1 and cyclin-dependent kinase 6 (CDK6) play essential roles in this process, although the exact mechanism remains to be investigated [137]. Nevertheless, in chronic hypoxia there is a HIF-1-independent increase in IκBα expression [138]. Additionally, IκBα inhibits HIF-1α hydroxylation by FIH [141], i.e., an inhibitor of the NF-κB pathway activates the HIFs activation pathway. In inflammatory reactions, proteolytic degradation of IκBα occurs, which inhibits HIF activation by increasing FIH activity [141]. Another mechanism that reduces inflammatory responses is the increase in PHD3 expression, which blocks the interaction of IKKβ and heat shock protein 90 (Hsp90), preventing the activation of this kinase [35,36,37,142]. On the other hand, Hsp90 induces activation of PRKD2 in chronic hypoxia, which activates NF-κB [126]. The two discussed pathways may interact via other mechanisms. For example, NF-κB increases in PHD3 expression [128]. This protein, independent of its enzymatic properties, is a coactivator of p65 NF-κB [128], similar to PHD2 [127]. PHD2 and PHD3 also reduce HIF activity by participating in the degradation of HIF-α subunits. Another mechanism is the involvement of HIF-1β in TNF receptor associated factor 6 (TRAF6) expression and therefore in NF-κB activation [143]. HIF-1α decreases TRAF6 expression, probably by binding HIF-1β into the HIF-1 complex.
Thanks to the aforementioned mechanisms, there is no simultaneous response of the cell to hypoxia and to pro-inflammatory factors. Nevertheless, NF-κB is activated in chronic hypoxia, leading to an increase in the expression of some inflammatory genes [40].

2.7. Chronic Hypoxia vs. Cycling Hypoxia in a Tumor

The signaling pathways activated during chronic hypoxia are very well understood. Hydroxylation of HIF-α is reduced, which results in an accumulation of these subunits in the cell. Phosphorylation by MAPK kinase, change in acetylation, or influence of ROS are also responsible for the increase in HIF-α stability during chronic hypoxia. There is also an activation of NF-κB, which increases the expression of HIF-1α. Ultimately, chronic hypoxia occurs in 23 to 54% of the tumor area, depending on the tumor model and the adoption of threshold oxygen levels from which hypoxia is defined [144,145]. In comparison, cycling hypoxia covers between 29 and 62% of the tumor area, depending on the tumor model and the adopted oxygen threshold level in the definition of hypoxia [144,145]. Nevertheless, this microenvironment is not often studied. For this reason, the activated signaling pathways and thus the cellular response to cycling hypoxia are poorly understood.

3. Cycling Hypoxia

3.1. Cycling Hypoxia in a Tumor

In the initial stages of tumor growth, the intense proliferation of tumor cells is not matched by the development of blood vessels that supply cells inside the tumor with nutrients and oxygen. Therefore, chronic (continuous, non-interrupted) hypoxia occurs inside the tumor [146]. This causes activation of signaling pathways that result in angiogenesis. The blood vessels produced in the tumor are characterized by structural abnormalities [147,148]. They do not show a conventional hierarchy compared to normal vessels, which impedes blood flow. In addition, endothelial cells and pericytes are poorly connected to each other, resulting in the leakiness of blood vessels in a tumor [149]. The structural abnormalities of blood vessels result in periodic oxygen deficiencies coupled with reoxygenation in various regions of the tumor [150]. This process is known as cycling (intermittent, transient) hypoxia. This is associated with changes in the vascular blood flow pathway characterized by the absence of conventional hierarchy. There is a segmentation of the tumor into regions that experience hypoxia and normoxia at a specific time [151]. Fluctuations in oxygen concentration range from a few minutes [144,145,151,152,153] to a few hours [154,155]. At the same time, the pattern of fluctuation depends on the type of tumor, including the line that produced the tumor in in vivo studies [151,154,156]. It has been shown that the more frequent the fluctuations in oxygen levels, the stronger the responses of the cells [157,158]. In contrast, the amplitude depends on the size of the tumor. The larger the tumor, the greater the fluctuations in oxygen concentration [156].
Cycling hypoxia is a characteristic feature of malignant tumors. This type of hypoxia is also associated with further tumor growth. In particular, cycling hypoxia increases the tumor growth rate [159,160]. It also causes apoptotic resistance by increasing B-cell lymphoma-extra large (Bcl-xL) expression in cancer cells [161]. Simultaneously, cycling hypoxia causes migration and metastases [162,163,164] associated with the induction of the epithelial-to-mesenchymal transition (EMT) of tumor cells. Additionally, cycling hypoxia increases self-renewal of cancer stem cells [163,164], which is associated with an increased expression of transcription factor BTB and CNC homology 1 (Bach1) [160]. In experiments on macrophages, cycling hypoxia increased the pro-inflammatory phenotype of M1 macrophages, while it had no pro-inflammatory effect on M2 macrophages [165]. On the other hand, experiments on lung carcinoma LLC1 cells have shown that cycling hypoxia reduces the number of M1 macrophages in a tumor [159]. In contrast, chronic hypoxia causes M2 polarization of macrophages [166].

3.2. Cycling Hypoxia: Intracellular Signaling Pathways

Cycling hypoxia alters the expression of fewer genes than chronic hypoxia [167], although the cellular response is similar for both. Cycling hypoxia has been shown to strongly activate the epidermal growth factor (EGF) pathway through a greater (compared to chronic hypoxia) increase in the expression of activators of the epidermal growth factor receptor (EGFR) family of receptors [167]. Cycling hypoxia, just like chronic hypoxia, is pro-inflammatory [168,169]. By activating NF-κB, it increases the expression of pro-inflammatory genes including cyclooxygenase-2 (COX-2).
Both discussed types of hypoxia alter the expression of similar genes, due to the fact that cycling hypoxia activates HIF-1 and NF-κB, the same transcription factors as chronic hypoxia [167,170]. However, the mechanisms of activation, as well as the degree of their activation, are different. In cycling hypoxia, activation of HIF-1 is stronger and longer [163,171,172]. In addition, the expression level of HIF-1α protein is increasingly higher with successive hypoxia cycles [173]. The higher the frequency of cycles (number of cycles per hour), the higher the activation of HIFs [158]. In contrast, reoxygenation is followed by HIF-α degradation, in part through increased PHD expression via HIF-1 activated in the period of reduced oxygen concentration [35,36,37,38]. The frequency of oxygen concentration fluctuations in the tumor depends on the cell line that produced the tumor [151,154,156]. Additionally, the more frequent the fluctuations in oxygen concentration, the greater the activation of NF-κB [157,158].
The mechanism of HIF-α accumulation in cycling hypoxia is ROS-dependent [161,174]. Cycling hypoxia induces the upregulation of NOX1 [173] and NOX4 [172,175], which generate ROS. ROS is also produced via the activation of xanthine oxidase [176,177], which may occur due to a NOX2-induced increase in intracellular calcium levels [177]. Another source of ROS in cycling hypoxia is the mitochondrial electron transport chain [178,179]. Increased levels of ROS increase the synthesis and stability of HIF-1α by decreasing PHD activity [171,180], probably due to the oxidation of the iron atom which is important in PHD activity [62,64,65]. ROS also inactivates FIH, although the exact mechanism of this inactivation has been poorly researched [31]. On the other hand, ROS causes activation of calpains, which cause HIF-2α degradation in cycling hypoxia [176]. Nevertheless, more research is required on whether HIF-2 has some role in cycling hypoxia.
Other signaling pathways also play an important role in cycling hypoxia. In particular, ROS activates MAPK cascades—ERK MAPK [175]—and the activation of JNK MAPK cascade leads to the activation of AP-1 [178]. However, there are no studies showing the effect of activated MAPK cascades in cycling hypoxia on phosphorylation of the HIF-1α subunit. This subunit may undergo phosphorylation in chronic hypoxia, which increases its stability and accumulation in the cell nucleus [92,95,181,182]. In cycling hypoxia, there is increased expression of DUSP1 [183,184]—a phosphatase that inactivates MAP kinases [103] but also increases the expression of manganese superoxide dismutase (MnSOD) that decreases ROS levels [184]. ROS activate phospholipase C-γ (PLCγ), resulting in increased calcium ion levels in the cytoplasm and PKC activation [177,180]. This leads to the activation of the mammalian target of rapamycin (mTOR). This kinase, probably through S6K, causes phosphorylation of HIF-1α and thus increases its stability.
Cycling hypoxia also causes activation of protein kinase A (PKA) (Figure 5) [185]. This kinase causes the phosphorylation of Thr63 and Ser692 in HIF-1α, which increases the stability of this HIF-1 subunit [185,186]. The mechanism of PKA activation is independent of cyclic adenosine monophosphate (cAMP) but is ROS-dependent [187]. The increase in HIF-1α stability induced by phosphorylation by PKA and mTOR is independent of PHD and oxygen levels.
In cycling hypoxia, ROS activates nuclear factor erythroid 2-related factor (Nrf2) [173], which results in an increased expression of thioredoxin 1 (Trx1), which then increases HIF-1α signaling. This effect is related to the interaction of Trx1 with HIF-1 in the cell nucleus, but not to the reductase activity of Trx1 [188,189].
Cycling hypoxia also causes changes in HIF-1α acetylation. Cycling hypoxia results in decreased expression of HDAC3 and HDAC5 proteins, but not the other HDACs [190], as demonstrated in rat pheochromocytoma PC12 cells. A change in HDAC5 activity levels increases HIF-1α acetylation and thus the stability of this HIF-1 subunit. With this model, HDAC3 does not appear to affect transcriptional activity or HIF-1α stability. There are no other (i.e., other than Wang et al. 2021) studies on the alteration of HIF-1α acetylation in cycling hypoxia. However, in normoxia and chronic hypoxia, this post-translational modification of the HIF-1α subunit is frequently studied [48,49,52,53,56].
In cycling hypoxia, ROS activates NF-κB [161,174,175,191]. In breast cancer, this is dependent on IκBα degradation [169]. In melanoma, this process is independent of IKK activation and of IκBα degradation [192]. NF-κB activation may depend on ERK MAPK, p38 MAPK, and JNK MAPK [170,179,193,194]. On the other hand, a study on cycling hypoxia in the renal tubular epithelial cell model showed a reduction in the expression of ubiquitin-specific peptidase 8 (USP8) [195], which leads to ubiquitination of Lys63 TAK1 and activation of the entire NF-κB activation pathway.

3.3. Cycling Hypoxia: Effects on the Tumor Microenvironment

Chronic hypoxia is accompanied by an activation of NF-κB, which increases the expression of HIF-1α and some pro-inflammatory genes [40]. However, it seems that, in cycling hypoxia, the activation of NF-κB is greater than in chronic hypoxia, and thus this type of hypoxia has a very pro-inflammatory character [168,169,192,196]. It increases the expression of genes associated with inflammatory responses as well as increases the cellular response to pro-inflammatory factors. In particular, there is an increased expression of COX-2 [167,168,169,197], CC motif chemokine ligand 2 (CCL2)/monocyte chemoattractant protein 1 (MCP-1) [192,194,198,199], CXC motif chemokine ligand (CXCL)1/growth related oncogene (GRO)-α [167], CXCL8/interleukin (IL)-8 [167,168,200,201], and IL-6 [168]. All of these are inflammatory mediators involved in various neoplastic processes.
Both types of hypoxia also increase vascular endothelial growth factor (VEGF)-A expression. This effect depends on the cancer cell line. VEGF-A expression in the tumor cell is increased much more under chronic hypoxic conditions than in cycling hypoxia. This has been shown in melanoma WM793B cells and prostate cancer PC-3 cells [167], as well as hepatocellular carcinoma HepG2 cells [202]. At the same time, in ovarian cancer SK-OV-3 cells, cycling hypoxia did not increase VEGF-A expression [167]. However, in melanoma A-07 cells, both types of hypoxia increased VEGF-A equally [203].
VEGF-A is one of the best described pro-angiogenic factors in a tumor (Figure 6) [204,205]. However, in a tumor, there is not only VEGF-A, but also other angiogenesis-inducing factors. These include factors whose expression is associated with cycling hypoxia, in particular the aforementioned CCL2/MCP-1, CXCL1/GRO-α, CXCL8/IL-8, and prostaglandin E2 (PGE2)—the product of COX-2 activity. The main mechanism of the proangiogenic properties of CCL2/MCP-1 is the recruitment of monocytes into the tumor niche, which are transformed into TAM [206,207], which secrete VEGF-A but also other proangiogenic factors such as matrix metalloproteinase 9 (MMP-9) and PGE2 [208]. CCL2/MCP-1 can also directly act on endothelial cells [209]. CXCL1/GRO-α and CXCL8/IL-8 are chemokines that activate CXC motif chemokine receptor (CXCR)2 [210,211], responsible for their pro-angiogenic properties [212,213,214]. These chemokines also recruit tumor-associated neutrophils (TAN) to the tumor niche [215,216,217,218]—cells that secrete MMP-9 into the tumor microenvironment [219,220]; MMP-9 is a metalloproteinase that causes a VEGF release from the extracellular matrix (ECM) [221].
PGE2 is also a pro-angiogenic factor, although not directly. It participates in angiogenesis and lymphangiogenesis by increasing the expression of various angiogenic and lymphangiogenic factors such as VEGF-A, VEGF-C, basic fibroblast growth factor (bFGF), platelet-derived growth factor (PDGF), endothelin-1 [222,223,224,225,226] and causes an increase in the expression of CXCR4, the receptor for angiogenic CXCL12 [227].
The aforementioned pro-inflammatory factors induced by cycling hypoxia also act on tumor-associated cells. For example, they recruit various cells into the tumor niche. CCL2/MCP-1 is a TAM recruiting factor [206,228,229], while CXCL1/GRO-α and CXCL8/IL-8 are TAN recruiting factors [215,216,217,218]. PGE2, through its action on anti-tumor cells, is one of the mechanisms of cancer immunoevasion. It inhibits the anticancer function of NK cells and dendritic cells and enhances the pro-cancer function of M2 macrophages and regulatory T cells (Treg) [230,231,232,233,234].

4. Mediators of Inflammatory Responses Induced by Chronic Hypoxia as a Therapeutic Target

Cycling hypoxia is a feature of all solid tumors [145,151,152,153,154,155]. It activates the same signaling pathways and alters the tumor microenvironment identically or similarly in all tumors. Therefore, understanding the mechanisms of action of cycling hypoxia will either allow the development or improvement of anti-cancer therapies against many types of cancer.
Cycling hypoxia is associated with elevated COX-2 expression and consequently an increase in PGE2 production [167,168,169,197]. For this reason, the use of nonsteroidal anti-inflammatory drugs (NSAID) together with standard anticancer therapy provides beneficial effects for patients with various solid tumors [235], especially breast cancer [236], colorectal cancer [235], oesophageal cancer [235], and prostate cancer [237]. Nevertheless, in patients with non-small-cell lung cancer, NSAIDs improve the overall response rate but have no effect on patient survival after therapy [238,239,240].
Additionally, cycling hypoxia increases CCL2/MCP-1 production in the tumor [192,194,198,199]. Therefore, taking the CCL2→CC motif chemokine receptor 2 (CCR2) axis as a therapeutic target is an approach with great therapeutic potential. In particular, a CCR2 antagonist [241,242,243] and CCL2-neutralizing antibody [244,245,246,247] are being tested for the treatment of many types of cancer.
In addition to CCL2/MCP-1, cycling hypoxia increases in the expression of CXCL1/GRO-α [167] and CXCL8/IL-8 [167,168,200,201]. For this reason, a CXCL1-neutralizing antibody [248] and CXCL8-neutralizing antibody [249,250,251,252] are being tested as potential anticancer agents. Another therapeutic approach is the use of receptor antagonists for sub-family CXC chemokines, such as CXCR2 antagonists SB225002 [253,254,255] and SB265610 [256]. CXCR1/CXCR2 dual antagonists that act on both CXCL8/IL-8 receptors have also been tested [257,258,259,260,261]. Because CXCR2 is a receptor for CXCL1/GRO-α [210], such dual antagonists also reduce the effects of this chemokine. It is also possible during cancer therapy to inhibit the entire NF-κB activation pathway by using proteasome inhibitors and IKKβ inhibitors [262]. This prevents the activation of NF-κB and so the expression of all genes is dependent on the activation of this transcription factor by cycling hypoxia.
Another option is to improve the anti-cancer anti-angiogenic therapy, e.g., by using bevacizumab—an anti-VEGF-A monoclonal antibody [263]. However, resistance to bevacizumab is very common, which is related to the presence of pro-angiogenic factors other than VEGF-A in the tumor. These factors complement or, in the absence of VEGF-A, replace VEGF-A in their functions [263,264]. For this reason, it has been suggested that bevacizumab be used together with drugs that inhibit other pro-angiogenic factors, particularly those induced by cycling hypoxia, such as anti-CCL2 antibody [265] and CCR2 inhibitor [266]. CCR2 is the receptor for CCL2/MCP-1 and both therapeutic approaches target the CCL2→CCR2 axis. Another option is to use bevacizumab with NSAID [267], mainly COX-2 inhibitors that reduce PGE2 production. As already mentioned, PGE2 has no direct angiogenic effect, but it increases the expression of pro-angiogenic factors [222,223,224,225,226]. Therefore, decreased PGE2 production results in decreased expression of other pro-angiogenic factors. Another possibility is to combine bevacizumab with a CXCR1/CXCR2 dual inhibitor [268]. It is also possible to combine bevacizumab with an inhibitor of the NF-κB activation pathway, e.g., NPI-0052/salinosporamide A, which is a proteasome inhibitor that blocks proteolytic degradation of IκBα [269]. NF-κB activation in cycling hypoxia is the most important mechanism in increasing the expression of all the aforementioned pro-angiogenic factors [168,169,192,196]. Therefore, decreased NF-κB activation decreases the expression of pro-angiogenic factors induced by this transcription factor.

5. Conclusions: A Perspective for Further Research on Chronic Hypoxia

The vast majority of published in vitro experiments on hypoxia in cancer relate to chronic hypoxia. Most of the available work has not investigated the effect of cyclic changes in oxygen concentration on tumor cells. For this reason, this type of research model does not reflect the actual state of the cancerous tumor, with cycling hypoxia affecting a considerable part of the tumor. In this way, the results of studies showing the effect of chronic hypoxia only reflect the situation in one area in a tumor. For this reason, it is advisable that each study on hypoxia in a tumor should use an in vitro model that includes cycling hypoxia.

Author Contributions

J.K., writing—original draft preparation; D.S., investigation; M.G.-D., investigation; J.L., investigation; I.G., investigation; D.C., funding acquisition, supervision; I.B.-B., writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the statutory budget of the Department of Biochemistry and Medical Chemistry Pomeranian Medical University in Szczecin, Poland.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chanmee, T.; Ontong, P.; Konno, K.; Itano, N. Tumor-associated macrophages as major players in the tumor microenvironment. Cancers 2014, 6, 1670–1690. [Google Scholar] [CrossRef] [Green Version]
  2. Najafi, M.; Farhood, B.; Mortezaee, K. Contribution of regulatory T cells to cancer: A review. J. Cell Physiol. 2019, 234, 7983–7993. [Google Scholar] [CrossRef]
  3. Ligęza, J.; Ligęza, J.; Klein, A. Growth factor/growth factor receptor loops in autocrine growth regulation of human prostate cancer DU145 cells. Acta Biochim. Pol. 2011, 58, 391–396. [Google Scholar] [CrossRef]
  4. Sulciner, M.L.; Gartung, A.; Gilligan, M.M.; Serhan, C.N.; Panigrahy, D. Targeting lipid mediators in cancer biology. Cancer Metastasis Rev. 2018, 37, 557–572. [Google Scholar] [CrossRef]
  5. Do, H.T.T.; Lee, C.H.; Cho, J. Chemokines and their Receptors: Multifaceted Roles in Cancer Progression and Potential Value as Cancer Prognostic Markers. Cancers 2020, 12, 287. [Google Scholar] [CrossRef] [Green Version]
  6. Dhup, S.; Dadhich, R.K.; Porporato, P.E.; Sonveaux, P. Multiple biological activities of lactic acid in cancer: Influences on tumor growth, angiogenesis and metastasis. Curr. Pharm. Des. 2012, 18, 1319–1330. [Google Scholar] [CrossRef] [Green Version]
  7. Medzhitov, R. Origin and physiological roles of inflammation. Nature 2008, 454, 428–435. [Google Scholar] [CrossRef]
  8. Muller, A.J.; Sharma, M.D.; Chandler, P.R.; Duhadaway, J.B.; Everhart, M.E.; Johnson, B.A., 3rd.; Kahler, D.J.; Pihkala, J.; Soler, A.P.; Munn, D.H.; et al. Chronic inflammation that facilitates tumor progression creates local immune suppression by inducing indoleamine 2,3 dioxygenase. Proc. Natl. Acad. Sci. USA 2008, 105, 17073–17078. [Google Scholar] [CrossRef] [Green Version]
  9. Chai, E.Z.; Siveen, K.S.; Shanmugam, M.K.; Arfuso, F.; Sethi, G. Analysis of the intricate relationship between chronic inflammation and cancer. Biochem. J. 2015, 468, 1–15. [Google Scholar] [CrossRef]
  10. Korbecki, J.; Kojder, K.; Barczak, K.; Simińska, D.; Gutowska, I.; Chlubek, D.; Baranowska-Bosiacka, I. Hypoxia Alters the Expression of CC Chemokines and CC Chemokine Receptors in a Tumor-A Literature Review. Int. J. Mol. Sci. 2020, 21, 5647. [Google Scholar] [CrossRef]
  11. Korbecki, J.; Kojder, K.; Kapczuk, P.; Kupnicka, P.; Gawrońska-Szklarz, B.; Gutowska, I.; Chlubek, D.; Baranowska-Bosiacka, I. The Effect of Hypoxia on the Expression of CXC Chemokines and CXC Chemokine Receptors-A Review of Literature. Int. J. Mol. Sci. 2021, 22, 843. [Google Scholar] [CrossRef]
  12. Tanaka, T.; Wiesener, M.; Bernhardt, W.; Eckardt, K.U.; Warnecke, C. The human HIF (hypoxia-inducible factor)-3alpha gene is a HIF-1 target gene and may modulate hypoxic gene induction. Biochem. J. 2009, 424, 143–151. [Google Scholar] [CrossRef] [Green Version]
  13. Zhang, P.; Yao, Q.; Lu, L.; Li, Y.; Chen, P.J.; Duan, C. Hypoxia-inducible factor 3 is an oxygen-dependent transcription activator and regulates a distinct transcriptional response to hypoxia. Cell Rep. 2014, 6, 1110–1121. [Google Scholar] [CrossRef] [Green Version]
  14. Yang, S.L.; Wu, C.; Xiong, Z.F.; Fang, X. Progress on hypoxia-inducible factor-3: Its structure, gene regulation and biological function (Review). Mol. Med. Rep. 2015, 12, 2411–2416. [Google Scholar] [CrossRef] [Green Version]
  15. Wu, C.; Yang, T.; Liu, Y.; Lu, Y.; Yang, Y.; Liu, X.; Liu, X.; Ye, L.; Sun, Y.; Wang, X.; et al. ARNT/HIF-1β links high-risk 1q21 gain and microenvironmental hypoxia to drug resistance and poor prognosis in multiple myeloma. Cancer Med. 2018, 7, 3899–3911. [Google Scholar] [CrossRef]
  16. Lee, J.S.; Kim, E.Y.; Iwabuchi, K.; Iwata, H. Molecular and functional characterization of aryl hydrocarbon receptor nuclear translocator 1 (ARNT1) and ARNT2 in chicken (Gallus gallus). Comp. Biochem. Physiol. C Toxicol. Pharm. 2011, 153, 269–279. [Google Scholar] [CrossRef]
  17. Kimura, Y.; Kasamatsu, A.; Nakashima, D.; Yamatoji, M.; Minakawa, Y.; Koike, K.; Fushimi, K.; Higo, M.; Endo-Sakamoto, Y.; Shiiba, M.; et al. ARNT2 Regulates Tumoral Growth in Oral Squamous Cell Carcinoma. J. Cancer 2016, 7, 702–710. [Google Scholar] [CrossRef] [Green Version]
  18. Yang, B.; Yang, E.; Liao, H.; Wang, Z.; Den, Z.; Ren, H. ARNT2 is downregulated and serves as a potential tumor suppressor gene in non-small cell lung cancer. Tumour Biol. 2015, 36, 2111–2119. [Google Scholar] [CrossRef]
  19. Li, W.; Liang, Y.; Yang, B.; Sun, H.; Wu, W. Downregulation of ARNT2 promotes tumor growth and predicts poor prognosis in human hepatocellular carcinoma. J. Gastroenterol. Hepatol. 2015, 30, 1085–1093. [Google Scholar] [CrossRef]
  20. Bruick, R.K.; McKnight, S.L. A conserved family of prolyl-4-hydroxylases that modify HIF. Science 2001, 294, 1337–1340. [Google Scholar] [CrossRef] [Green Version]
  21. Landázuri, M.O.; Vara-Vega, A.; Vitón, M.; Cuevas, Y.; del Peso, L. Analysis of HIF-prolyl hydroxylases binding to substrates. Biochem. Biophys. Res. Commun. 2006, 351, 313–320. [Google Scholar] [CrossRef]
  22. 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]
  23. Tuckerman, J.R.; Zhao, Y.; Hewitson, K.S.; Tian, Y.M.; Pugh, C.W.; Ratcliffe, P.J.; Mole, D.R. Determination and comparison of specific activity of the HIF-prolyl hydroxylases. FEBS Lett. 2004, 576, 145–150. [Google Scholar] [CrossRef] [Green Version]
  24. Masson, N.; Willam, C.; Maxwell, P.H.; Pugh, C.W.; Ratcliffe, P.J. Independent function of two destruction domains in hypoxia-inducible factor-alpha chains activated by prolyl hydroxylation. EMBO J. 2001, 20, 5197–5206. [Google Scholar] [CrossRef] [Green Version]
  25. Maxwell, P.H.; Wiesener, M.S.; Chang, G.W.; Clifford, S.C.; Vaux, E.C.; Cockman, M.E.; Wykoff, C.C.; Pugh, C.W.; Maher, E.R.; Ratcliffe, P.J. The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature 1999, 399, 271–275. [Google Scholar] [CrossRef]
  26. 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]
  27. Hon, W.C.; Wilson, M.I.; Harlos, K.; Claridge, T.D.; Schofield, C.J.; Pugh, C.W.; Maxwell, P.H.; Ratcliffe, P.J.; Stuart, D.I.; Jones, E.Y. Structural basis for the recognition of hydroxyproline in HIF-1 alpha by pVHL. Nature 2002, 417, 975–978. [Google Scholar] [CrossRef]
  28. Cockman, M.E.; Masson, N.; Mole, D.R.; Jaakkola, P.; Chang, G.W.; Clifford, S.C.; Maher, E.R.; Pugh, C.W.; Ratcliffe, P.J.; Maxwell, P.H. Hypoxia inducible factor-alpha binding and ubiquitylation by the von Hippel-Lindau tumor suppressor protein. J. Biol. Chem. 2000, 275, 25733–25741. [Google Scholar] [CrossRef] [Green Version]
  29. Corn, P.G.; McDonald, E.R., 3rd.; Herman, J.G.; El-Deiry, W.S. Tat-binding protein-1, a component of the 26S proteasome, contributes to the E3 ubiquitin ligase function of the von Hippel-Lindau protein. Nat. Genet. 2003, 35, 229–237. [Google Scholar] [CrossRef]
  30. Lando, D.; Peet, D.J.; Whelan, D.A.; Gorman, J.J.; Whitelaw, M.L. Asparagine hydroxylation of the HIF transactivation domain a hypoxic switch. Science 2002, 295, 858–861. [Google Scholar] [CrossRef]
  31. 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]
  32. Dames, S.A.; Martinez-Yamout, M.; De Guzman, R.N.; Dyson, H.J.; Wright, P.E. Structural basis for Hif-1 alpha/CBP recognition in the cellular hypoxic response. Proc. Natl. Acad. Sci. USA 2002, 99, 5271–5276. [Google Scholar] [CrossRef] [Green Version]
  33. Freedman, S.J.; Sun, Z.Y.; Poy, F.; Kung, A.L.; Livingston, D.M.; Wagner, G.; Eck, M.J. Structural basis for recruitment of CBP/p300 by hypoxia-inducible factor-1 alpha. Proc. Natl. Acad. Sci. USA 2002, 99, 5367–5372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Hirsilä, M.; Koivunen, P.; Günzler, V.; Kivirikko, K.I.; Myllyharju, J. Characterization of the human prolyl 4-hydroxylases that modify the hypoxia-inducible factor. J. Biol. Chem. 2003, 278, 30772–30780. [Google Scholar] [CrossRef] [Green Version]
  35. D’Angelo, G.; Duplan, E.; Boyer, N.; Vigne, P.; Frelin, C. Hypoxia up-regulates prolyl hydroxylase activity: A feedback mechanism that limits HIF-1 responses during reoxygenation. J. Biol. Chem. 2003, 278, 38183–38187. [Google Scholar] [CrossRef] [Green Version]
  36. Stiehl, D.P.; Wirthner, R.; Köditz, J.; Spielmann, P.; Camenisch, G.; Wenger, R.H. Increased prolyl 4-hydroxylase domain proteins compensate for decreased oxygen levels. Evidence for an autoregulatory oxygen-sensing system. J. Biol. Chem. 2006, 281, 23482–23491. [Google Scholar] [CrossRef] [Green Version]
  37. Ginouvès, A.; Ilc, K.; Macías, N.; Pouysségur, J.; Berra, E. PHDs overactivation during chronic hypoxia “desensitizes” HIFalpha and protects cells from necrosis. Proc. Natl. Acad. Sci. USA 2008, 105, 4745–4750. [Google Scholar] [CrossRef] [Green Version]
  38. Fujita, N.; Markova, D.; Anderson, D.G.; Chiba, K.; Toyama, Y.; Shapiro, I.M.; Risbud, M.V. Expression of prolyl hydroxylases (PHDs) is selectively controlled by HIF-1 and HIF-2 proteins in nucleus pulposus cells of the intervertebral disc: Distinct roles of PHD2 and PHD3 proteins in controlling HIF-1α activity in hypoxia. J. Biol. Chem. 2012, 287, 16975–16986. [Google Scholar] [CrossRef] [Green Version]
  39. Koivunen, P.; Hirsilä, M.; Günzler, 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] [Green Version]
  40. Ravenna, L.; Principessa, L.; Verdina, A.; Salvatori, L.; Russo, M.A.; Petrangeli, E. Distinct phenotypes of human prostate cancer cells associate with different adaptation to hypoxia and pro-inflammatory gene expression. PLoS ONE 2014, 9, e96250. [Google Scholar] [CrossRef]
  41. Guan, Z.; Ding, C.; Du, Y.; Zhang, K.; Zhu, J.N.; Zhang, T.; He, D.; Xu, S.; Wang, X.; Fan, J. HAF drives the switch of HIF-1α to HIF-2α by activating the NF-κB pathway, leading to malignant behavior of T24 bladder cancer cells. Int. J. Oncol. 2014, 44, 393–402. [Google Scholar] [CrossRef] [Green Version]
  42. Luo, W.; Zhong, J.; Chang, R.; Hu, H.; Pandey, A.; Semenza, G.L. Hsp70 and CHIP selectively mediate ubiquitination and degradation of hypoxia-inducible factor (HIF)-1alpha but Not HIF-2alpha. J. Biol. Chem. 2010, 285, 3651–3663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Munksgaard Persson, M.; Johansson, M.E.; Monsef, N.; Planck, M.; Beckman, S.; Seckl, M.J.; Rönnstrand, L.; Påhlman, S.; Pettersson, H.M. HIF-2α expression is suppressed in SCLC cells, which survive in moderate and severe hypoxia when HIF-1α is repressed. Am. J. Pathol. 2012, 180, 494–504. [Google Scholar] [CrossRef] [Green Version]
  44. 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]
  45. Imtiyaz, H.Z.; Williams, E.P.; Hickey, M.M.; Patel, S.A.; Durham, A.C.; Yuan, L.J.; Hammond, R.; Gimotty, P.A.; Keith, B.; Simon, M.C. Hypoxia-inducible factor 2alpha regulates macrophage function in mouse models of acute and tumor inflammation. J. Clin. Invest. 2010, 120, 2699–2714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Burke, B.; Tang, N.; Corke, K.P.; Tazzyman, D.; Ameri, K.; Wells, M.; Lewis, C.E. Expression of HIF-1alpha by human macrophages: Implications for the use of macrophages in hypoxia-regulated cancer gene therapy. J. Pathol. 2002, 196, 204–212. [Google Scholar] [CrossRef]
  47. Kim, W.; Bennett, E.J.; Huttlin, E.L.; Guo, A.; Li, J.; Possemato, A.; Sowa, M.E.; Rad, R.; Rush, J.; Comb, M.J.; et al. Systematic and quantitative assessment of the ubiquitin-modified proteome. Mol. Cell 2011, 44, 325–340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Tang, Y.; Liu, S.; Li, N.; Guo, W.; Shi, J.; Yu, H.; Zhang, L.; Wang, K.; Liu, S.; Cheng, S. 14–3-3ζ promotes hepatocellular carcinoma venous metastasis by modulating hypoxia-inducible factor-1α. Oncotarget 2016, 7, 15854–15867. [Google Scholar] [CrossRef] [Green Version]
  49. Kim, S.H.; Jeong, J.W.; Park, J.A.; Lee, J.W.; Seo, J.H.; Jung, B.K.; Bae, M.K.; Kim, K.W. Regulation of the HIF-1alpha stability by histone deacetylases. Oncol. Rep. 2007, 17, 647–651. [Google Scholar]
  50. Qian, D.Z.; Kachhap, S.K.; Collis, S.J.; Verheul, H.M.; Carducci, M.A.; Atadja, P.; Pili, R. Class II histone deacetylases are associated with VHL-independent regulation of hypoxia-inducible factor 1 alpha. Cancer Res. 2006, 66, 8814–8821. [Google Scholar] [CrossRef] [Green Version]
  51. Seo, H.W.; Kim, E.J.; Na, H.; Lee, M.O. Transcriptional activation of hypoxia-inducible factor-1alpha by HDAC4 and HDAC5 involves differential recruitment of p300 and FIH-1. FEBS Lett. 2009, 583, 55–60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Geng, H.; Harvey, C.T.; Pittsenbarger, J.; Liu, Q.; Beer, T.M.; Xue, C.; Qian, D.Z. HDAC4 protein regulates HIF1α protein lysine acetylation and cancer cell response to hypoxia. J. Biol. Chem. 2011, 286, 38095–38102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Kato, H.; Tamamizu-Kato, S.; Shibasaki, F. Histone deacetylase 7 associates with hypoxia-inducible factor 1alpha and increases transcriptional activity. J. Biol. Chem. 2004, 279, 41966–41974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Jeong, J.W.; Bae, M.K.; Ahn, M.Y.; Kim, S.H.; Sohn, T.K.; Bae, M.H.; Yoo, M.A.; Song, E.J.; Lee, K.J.; Kim, K.W. Regulation and destabilization of HIF-1alpha by ARD1-mediated acetylation. Cell 2002, 111, 709–720. [Google Scholar] [CrossRef] [Green Version]
  55. Geng, H.; Liu, Q.; Xue, C.; David, L.L.; Beer, T.M.; Thomas, G.V.; Dai, M.S.; Qian, D.Z. HIF1α protein stability is increased by acetylation at lysine 709. J. Biol. Chem. 2012, 287, 35496–35505. [Google Scholar] [CrossRef] [Green Version]
  56. Lim, J.H.; Lee, Y.M.; Chun, Y.S.; Chen, J.; Kim, J.E.; Park, J.W. Sirtuin 1 modulates cellular responses to hypoxia by deacetylating hypoxia-inducible factor 1alpha. Mol. Cell 2010, 38, 864–878. [Google Scholar] [CrossRef] [PubMed]
  57. Seo, K.S.; Park, J.H.; Heo, J.Y.; Jing, K.; Han, J.; Min, K.N.; Kim, C.; Koh, G.Y.; Lim, K.; Kang, G.Y.; et al. SIRT2 regulates tumour hypoxia response by promoting HIF-1α hydroxylation. Oncogene 2015, 34, 1354–1362. [Google Scholar] [CrossRef]
  58. Finley, L.W.; Carracedo, A.; Lee, J.; Souza, A.; Egia, A.; Zhang, J.; Teruya-Feldstein, J.; Moreira, P.I.; Cardoso, S.M.; Clish, C.B.; et al. SIRT3 opposes reprogramming of cancer cell metabolism through HIF1α destabilization. Cancer Cell 2011, 19, 416–428. [Google Scholar] [CrossRef] [Green Version]
  59. Hubbi, M.E.; Hu, H.; Kshitiz; Gilkes, D.M.; Semenza, G.L. Sirtuin-7 inhibits the activity of hypoxia-inducible factors. J. Biol. Chem. 2013, 288, 20768–20775. [Google Scholar] [CrossRef] [Green Version]
  60. 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] [Green Version]
  61. Kulisz, A.; Chen, N.; Chandel, N.S.; Shao, Z.; Schumacker, P.T. Mitochondrial ROS initiate phosphorylation of p38 MAP kinase during hypoxia in cardiomyocytes. Am. J. Physiol. Lung Cell. Mol. Physiol. 2002, 282, L1324–L1329. [Google Scholar] [CrossRef] [Green Version]
  62. 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]
  63. Tang, X.; Guo, D.; Lin, C.; Shi, Z.; Qian, R.; Fu, W.; Liu, J.; Li, X.; Fan, L. hCLOCK Causes Rho-Kinase-Mediated Endothelial Dysfunction and NF-κB-Mediated Inflammatory Responses. Oxid. Med. Cell Longev. 2015, 2015, 671839. [Google Scholar] [CrossRef] [Green Version]
  64. Köhl, R.; Zhou, J.; Brüne, B. Reactive oxygen species attenuate nitric-oxide-mediated hypoxia-inducible factor-1alpha stabilization. Free Radic. Biol. Med. 2006, 40, 1430–1442. [Google Scholar] [CrossRef] [PubMed]
  65. Gerald, D.; Berra, E.; Frapart, Y.M.; Chan, D.A.; Giaccia, A.J.; Mansuy, D.; Pouysségur, J.; Yaniv, M.; Mechta-Grigoriou, F. JunD reduces tumor angiogenesis by protecting cells from oxidative stress. Cell 2004, 118, 781–794. [Google Scholar] [CrossRef] [PubMed]
  66. Bonello, S.; Zähringer, C.; BelAiba, R.S.; Djordjevic, T.; Hess, J.; Michiels, C.; Kietzmann, T.; Görlach, A. Reactive oxygen species activate the HIF-1alpha promoter via a functional NFkappaB site. Arter. Thromb Vasc. Biol. 2007, 27, 755–761. [Google Scholar] [CrossRef] [Green Version]
  67. Morgan, M.J.; Liu, Z.G. Crosstalk of reactive oxygen species and NF-κB signaling. Cell Res. 2011, 21, 103–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Diebold, I.; Petry, A.; Hess, J.; Görlach, A. The NADPH oxidase subunit NOX4 is a new target gene of the hypoxia-inducible factor-1. Mol. Biol. Cell 2010, 21, 2087–2096. [Google Scholar] [CrossRef] [Green Version]
  69. Fitzgerald, J.P.; Nayak, B.; Shanmugasundaram, K.; Friedrichs, W.; Sudarshan, S.; Eid, A.A.; DeNapoli, T.; Parekh, D.J.; Gorin, Y.; Block, K. Nox4 mediates renal cell carcinoma cell invasion through hypoxia-induced interleukin 6- and 8- production. PLoS ONE 2012, 7, e30712. [Google Scholar] [CrossRef]
  70. Lu, X.; Murphy, T.C.; Nanes, M.S.; Hart, C.M. PPAR{gamma} regulates hypoxia-induced Nox4 expression in human pulmonary artery smooth muscle cells through NF-κB. Am. J. Physiol. Lung Cell. Mol. Physiol. 2010, 299, L559–L566. [Google Scholar] [CrossRef] [Green Version]
  71. 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]
  72. Aaltoma, S.H.; Lipponen, P.K.; Kosma, V.M. Inducible nitric oxide synthase (iNOS) expression and its prognostic value in prostate cancer. Anticancer Res. 2001, 21, 3101–3106. [Google Scholar] [PubMed]
  73. Bulut, A.S.; Erden, E.; Sak, S.D.; Doruk, H.; Kursun, N.; Dincol, D. Significance of inducible nitric oxide synthase expression in benign and malignant breast epithelium: An immunohistochemical study of 151 cases. Virchows Arch. 2005, 447, 24–30. [Google Scholar] [CrossRef] [PubMed]
  74. Ekmekcioglu, S.; Ellerhorst, J.A.; Prieto, V.G.; Johnson, M.M.; Broemeling, L.D.; Grimm, E.A. Tumor iNOS predicts poor survival for stage III melanoma patients. Int. J. Cancer 2006, 119, 861–866. [Google Scholar] [CrossRef] [PubMed]
  75. Li, F.; Sonveaux, P.; Rabbani, Z.N.; Liu, S.; Yan, B.; Huang, Q.; Vujaskovic, Z.; Dewhirst, M.W.; Li, C.Y. Regulation of HIF-1alpha stability through S-nitrosylation. Mol. Cell 2007, 26, 63–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. 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]
  77. Sogawa, K.; Numayama-Tsuruta, K.; Ema, M.; Abe, M.; Abe, H.; Fujii-Kuriyama, Y. Inhibition of hypoxia-inducible factor 1 activity by nitric oxide donors in hypoxia. Proc. Natl. Acad. Sci. USA 1998, 95, 7368–7373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Herr, B.; Zhou, J.; Dröse, S.; Brüne, B. The interaction of superoxide with nitric oxide destabilizes hypoxia-inducible factor-1alpha. Cell Mol. Life Sci. 2007, 64, 3295–3305. [Google Scholar] [CrossRef] [PubMed]
  79. Callapina, M.; Zhou, J.; Schmid, T.; Köhl, R.; Brüne, B. NO restores HIF-1alpha hydroxylation during hypoxia: Role of reactive oxygen species. Free Radic Biol. Med. 2005, 39, 925–936. [Google Scholar] [CrossRef]
  80. Salnikow, K.; Kluz, T.; Costa, M.; Piquemal, D.; Demidenko, Z.N.; Xie, K.; Blagosklonny, M.V. The regulation of hypoxic genes by calcium involves c-Jun/AP-1, which cooperates with hypoxia-inducible factor 1 in response to hypoxia. Mol. Cell Biol. 2002, 22, 1734–1741. [Google Scholar] [CrossRef] [Green Version]
  81. Yadav, S.; Kalra, N.; Ganju, L.; Singh, M. Activator protein-1 (AP-1): A bridge between life and death in lung epithelial (A549) cells under hypoxia. Mol. Cell Biochem. 2017, 436, 99–110. [Google Scholar] [CrossRef] [PubMed]
  82. Premkumar, D.R.; Adhikary, G.; Overholt, J.L.; Simonson, M.S.; Cherniack, N.S.; Prabhakar, N.R. Intracellular pathways linking hypoxia to activation of c-fos and AP-1. Adv. Exp. Med. Biol. 2000, 475, 101–109. [Google Scholar] [PubMed]
  83. Xu, L.; Pathak, P.S.; Fukumura, D. Hypoxia-induced activation of p38 mitogen-activated protein kinase and phosphatidylinositol 3′-kinase signaling pathways contributes to expression of interleukin 8 in human ovarian carcinoma cells. Clin. Cancer Res. 2004, 10, 701–707. [Google Scholar] [CrossRef] [Green Version]
  84. Lan, A.P.; Xiao, L.C.; Yang, Z.L.; Yang, C.T.; Wang, X.Y.; Chen, P.X.; Gu, M.F.; Feng, J.Q. Interaction between ROS and p38MAPK contributes to chemical hypoxia-induced injuries in PC12 cells. Mol. Med. Rep. 2012, 5, 250–255. [Google Scholar] [PubMed] [Green Version]
  85. Mottet, D.; Michel, G.; Renard, P.; Ninane, N.; Raes, M.; Michiels, C. ERK and calcium in activation of HIF-1. Ann. N. Y. Acad Sci. 2002, 973, 448–453. [Google Scholar] [CrossRef] [PubMed]
  86. Minet, E.; Michel, G.; Mottet, D.; Piret, J.P.; Barbieux, A.; Raes, M.; Michiels, C. c-JUN gene induction and AP-1 activity is regulated by a JNK-dependent pathway in hypoxic HepG2 cells. Exp. Cell Res. 2001, 265, 114–124. [Google Scholar] [CrossRef]
  87. Singh, M.; Yadav, S.; Kumar, M.; Saxena, S.; Saraswat, D.; Bansal, A.; Singh, S.B. The MAPK-activator protein-1 signaling regulates changes in lung tissue of rat exposed to hypobaric hypoxia. J. Cell Physiol. 2018, 233, 6851–6865. [Google Scholar] [CrossRef]
  88. Laderoute, K.R.; Calaoagan, J.M.; Gustafson-Brown, C.; Knapp, A.M.; Li, G.C.; Mendonca, H.L.; Ryan, H.E.; Wang, Z.; Johnson, R.S. The response of c-jun/AP-1 to chronic hypoxia is hypoxia-inducible factor 1 alpha dependent. Mol. Cell Biol. 2002, 22, 2515–2523. [Google Scholar] [CrossRef] [Green Version]
  89. Scortegagna, M.; Cataisson, C.; Martin, R.J.; Hicklin, D.J.; Schreiber, R.D.; Yuspa, S.H.; Arbeit, J.M. HIF-1alpha regulates epithelial inflammation by cell autonomous NFkappaB activation and paracrine stromal remodeling. Blood 2008, 111, 3343–3354. [Google Scholar] [CrossRef] [Green Version]
  90. Kwon, S.J.; Song, J.J.; Lee, Y.J. Signal pathway of hypoxia-inducible factor-1alpha phosphorylation and its interaction with von Hippel-Lindau tumor suppressor protein during ischemia in MiaPaCa-2 pancreatic cancer cells. Clin. Cancer Res. 2005, 11, 7607–7613. [Google Scholar] [CrossRef] [Green Version]
  91. Comerford, K.M.; Cummins, E.P.; Taylor, C.T. c-Jun NH2-terminal kinase activation contributes to hypoxia-inducible factor 1alpha-dependent P-glycoprotein expression in hypoxia. Cancer Res. 2004, 64, 9057–9061. [Google Scholar] [CrossRef] [Green Version]
  92. Mylonis, I.; Chachami, G.; Samiotaki, M.; Panayotou, G.; Paraskeva, E.; Kalousi, A.; Georgatsou, E.; Bonanou, S.; Simos, G. Identification of MAPK phosphorylation sites and their role in the localization and activity of hypoxia-inducible factor-1alpha. J. Biol. Chem. 2006, 281, 33095–33106. [Google Scholar] [CrossRef] [Green Version]
  93. Karapetsas, A.; Giannakakis, A.; Pavlaki, M.; Panayiotidis, M.; Sandaltzopoulos, R.; Galanis, A. Biochemical and molecular analysis of the interaction between ERK2 MAP kinase and hypoxia inducible factor-1α. Int. J. Biochem. Cell Biol. 2011, 43, 1582–1590. [Google Scholar] [CrossRef]
  94. Khurana, A.; Nakayama, K.; Williams, S.; Davis, R.J.; Mustelin, T.; Ronai, Z. Regulation of the ring finger E3 ligase Siah2 by p38 MAPK. J. Biol. Chem. 2006, 281, 35316–35326. [Google Scholar] [CrossRef] [Green Version]
  95. Liu, C.; Shi, Y.; Du, Y.; Ning, X.; Liu, N.; Huang, D.; Liang, J.; Xue, Y.; Fan, D. Dual-specificity phosphatase DUSP1 protects overactivation of hypoxia-inducible factor 1 through inactivating ERK MAPK. Exp. Cell Res. 2005, 309, 410–418. [Google Scholar] [CrossRef]
  96. Mishra, O.P.; Delivoria-Papadopoulos, M. Effect of hypoxia on the expression and activity of mitogen-activated protein (MAP) kinase-phosphatase-1 (MKP-1) and MKP-3 in neuronal nuclei of newborn piglets: The role of nitric oxide. Neuroscience 2004, 129, 665–673. [Google Scholar] [CrossRef]
  97. Short, M.D.; Fox, S.M.; Lam, C.F.; Stenmark, K.R.; Das, M. Protein kinase Czeta attenuates hypoxia-induced proliferation of fibroblasts by regulating MAP kinase phosphatase-1 expression. Mol. Biol. Cell 2006, 17, 1995–2008. [Google Scholar] [CrossRef] [Green Version]
  98. Seta, K.A.; Kim, R.; Kim, H.W.; Millhorn, D.E.; Beitner-Johnson, D. Hypoxia-induced regulation of MAPK phosphatase-1 as identified by subtractive suppression hybridization and cDNA microarray analysis. J. Biol. Chem. 2001, 276, 44405–44412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Li, Q.; Ke, Q.; Costa, M. Alterations of histone modifications by cobalt compounds. Carcinogenesis 2009, 30, 1243–1251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Lamadema, N.; Burr, S.; Brewer, A.C. Dynamic regulation of epigenetic demethylation by oxygen availability and cellular redox. Free Radic. Biol. Med. 2019, 131, 282–298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Lin, S.C.; Chien, C.W.; Lee, J.C.; Yeh, Y.C.; Hsu, K.F.; Lai, Y.Y.; Lin, S.C.; Tsai, S.J. Suppression of dual-specificity phosphatase-2 by hypoxia increases chemoresistance and malignancy in human cancer cells. J. Clin. Investig. 2011, 121, 1905–1916. [Google Scholar] [CrossRef]
  102. Wu, M.H.; Lin, S.C.; Hsiao, K.Y.; Tsai, S.J. Hypoxia-inhibited dual-specificity phosphatase-2 expression in endometriotic cells regulates cyclooxygenase-2 expression. J. Pathol. 2011, 225, 390–400. [Google Scholar] [CrossRef]
  103. Patterson, K.I.; Brummer, T.; O’Brien, P.M.; Daly, R.J. Dual-specificity phosphatases: Critical regulators with diverse cellular targets. Biochem. J. 2009, 418, 475–489. [Google Scholar] [CrossRef] [Green Version]
  104. Mills, C.N.; Joshi, S.S.; Niles, R.M. Expression and function of hypoxia inducible factor-1 alpha in human melanoma under non-hypoxic conditions. Mol. Cancer 2009, 8, 104. [Google Scholar] [CrossRef] [Green Version]
  105. 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]
  106. Secades, P.; de Santa-María, I.S.; Merlo, A.; Suarez, C.; Chiara, M.D. In vitro study of normoxic epidermal growth factor receptor-induced hypoxia-inducible factor-1-alpha, vascular endothelial growth factor, and BNIP3 expression in head and neck squamous cell carcinoma cell lines: Implications for anti-epidermal growth factor receptor therapy. Head Neck 2015, 37, 1150–1162. [Google Scholar]
  107. Zampetaki, A.; Mitsialis, S.A.; Pfeilschifter, J.; Kourembanas, S. Hypoxia induces macrophage inflammatory protein-2 (MIP-2) gene expression in murine macrophages via NF-kappaB: The prominent role of p42/ p44 and PI3 kinase pathways. FASEB J. 2004, 18, 1090–1092. [Google Scholar] [CrossRef] [PubMed]
  108. 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]
  109. Van Uden, P.; Kenneth, N.S.; Rocha, S. Regulation of hypoxia-inducible factor-1alpha by NF-kappaB. Biochem. J. 2008, 412, 477–484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Qiao, Q.; Nozaki, Y.; Sakoe, K.; Komatsu, N.; Kirito, K. NF-κB mediates aberrant activation of HIF-1 in malignant lymphoma. Exp. Hematol. 2010, 38, 1199–1208. [Google Scholar] [CrossRef] [PubMed]
  111. Zhang, Z.; Huang, Y.; Zhang, J.; Liu, Z.; Lin, Q.; Wang, Z. Activation of NF-κB signaling pathway during HCG-induced VEGF expression in luteal cells. Cell Biol. Int. 2019, 43, 344–349. [Google Scholar] [CrossRef] [PubMed]
  112. 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] [PubMed] [Green Version]
  113. Nam, S.Y.; Ko, Y.S.; Jung, J.; Yoon, J.; Kim, Y.H.; Choi, Y.J.; Park, J.W.; Chang, M.S.; Kim, W.H.; Lee, B.L. A hypoxia-dependent upregulation of hypoxia-inducible factor-1 by nuclear factor-κB promotes gastric tumour growth and angiogenesis. Br. J. Cancer 2011, 104, 166–174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. 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]
  115. Van Uden, P.; Kenneth, N.S.; Webster, R.; Müller, H.A.; Mudie, S.; Rocha, S. Evolutionary conserved regulation of HIF-1β by NF-κB. PLoS Genet. 2011, 7, e1001285. [Google Scholar] [CrossRef] [Green Version]
  116. Baldea, I.; Teacoe, I.; Olteanu, D.E.; Vaida-Voievod, C.; Clichici, A.; Sirbu, A.; Filip, G.A.; Clichici, S. Effects of different hypoxia degrees on endothelial cell cultures-Time course study. Mech. Ageing Dev. 2018, 172, 45–50. [Google Scholar] [CrossRef]
  117. Cummins, E.P.; Berra, E.; Comerford, K.M.; Ginouves, A.; Fitzgerald, K.T.; Seeballuck, F.; Godson, C.; Nielsen, J.E.; Moynagh, P.; Pouyssegur, J.; et al. Prolyl hydroxylase-1 negatively regulates IkappaB kinase-beta, giving insight into hypoxia-induced NFkappaB activity. Proc. Natl. Acad. Sci. USA 2006, 103, 18154–18159. [Google Scholar] [CrossRef] [Green Version]
  118. Fitzpatrick, S.F.; Fábián, Z.; Schaible, B.; Lenihan, C.R.; Schwarzl, T.; Rodriguez, J.; Zheng, X.; Li, Z.; Tambuwala, M.M.; Higgins, D.G.; et al. Prolyl hydroxylase-1 regulates hepatocyte apoptosis in an NF-κB-dependent manner. Biochem. Biophys. Res. Commun. 2016, 474, 579–586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Wang, Y.; Zhao, W.; Gao, Q.; Fan, L.; Qin, Y.; Zhou, H.; Li, M.; Fang, J. pVHL mediates K63-linked ubiquitination of IKKβ, leading to IKKβ inactivation. Cancer Lett. 2016, 383, 1–8. [Google Scholar] [CrossRef] [PubMed]
  120. Wang, L.; Niu, Z.; Wang, X.; Li, Z.; Liu, Y.; Luo, F.; Yan, X. PHD2 exerts anti-cancer and anti-inflammatory effects in colon cancer xenografts mice via attenuating NF-κB activity. Life Sci. 2020, 242, 117167. [Google Scholar] [CrossRef]
  121. Culver, C.; Sundqvist, A.; Mudie, S.; Melvin, A.; Xirodimas, D.; Rocha, S. Mechanism of hypoxia-induced NF-kappaB. Mol. Cell Biol. 2010, 30, 4901–4921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Devries, I.L.; Hampton-Smith, R.J.; Mulvihill, M.M.; Alverdi, V.; Peet, D.J.; Komives, E.A. Consequences of IkappaB alpha hydroxylation by the factor inhibiting HIF (FIH). FEBS Lett. 2010, 584, 4725–4730. [Google Scholar] [CrossRef] [PubMed]
  123. Hsieh, K.Y.; Wei, C.K.; Wu, C.C. YC-1 Prevents Tumor-Associated Tissue Factor Expression and Procoagulant Activity in Hypoxic Conditions by Inhibiting p38/NF-κB Signaling Pathway. Int. J. Mol. Sci. 2019, 20, 244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Tang, Y.; Lv, P.; Sun, Z.; Han, L.; Luo, B.; Zhou, W. 14–3-3ζ up-regulates hypoxia-inducible factor-1α in hepatocellular carcinoma via activation of PI3K/Akt/NF-кB signal transduction pathway. Int. J. Clin. Exp. Pathol. 2015, 8, 15845–15853. [Google Scholar]
  125. Li, Y.; Yang, L.; Dong, L.; Yang, Z.W.; Zhang, J.; Zhang, S.L.; Niu, M.J.; Xia, J.W.; Gong, Y.; Zhu, N.; et al. Crosstalk between the Akt/mTORC1 and NF-κB signaling pathways promotes hypoxia-induced pulmonary hypertension by increasing DPP4 expression in PASMCs. Acta Pharm. Sin. 2019, 40, 1322–1333. [Google Scholar] [CrossRef]
  126. Azoitei, N.; Diepold, K.; Brunner, C.; Rouhi, A.; Genze, F.; Becher, A.; Kestler, H.; van Lint, J.; Chiosis, G.; Koren, J., 3rd.; et al. HSP90 supports tumor growth and angiogenesis through PRKD2 protein stabilization. Cancer Res. 2014, 74, 7125–7136. [Google Scholar] [CrossRef] [Green Version]
  127. Li, J.; Yuan, W.; Jiang, S.; Ye, W.; Yang, H.; Shapiro, I.M.; Risbud, M.V. Prolyl-4-hydroxylase domain protein 2 controls NF-κB/p65 transactivation and enhances the catabolic effects of inflammatory cytokines on cells of the nucleus pulposus. J. Biol. Chem. 2015, 290, 7195–7207. [Google Scholar] [CrossRef] [Green Version]
  128. Fujita, N.; Gogate, S.S.; Chiba, K.; Toyama, Y.; Shapiro, I.M.; Risbud, M.V. Prolyl hydroxylase 3 (PHD3) modulates catabolic effects of tumor necrosis factor-α (TNF-α) on cells of the nucleus pulposus through co-activation of nuclear factor κB (NF-κB)/p65 signaling. J. Biol. Chem. 2012, 287, 39942–39953. [Google Scholar] [CrossRef] [Green Version]
  129. Fang, H.Y.; Hughes, R.; Murdoch, C.; Coffelt, S.B.; Biswas, S.K.; Harris, A.L.; Johnson, R.S.; Imityaz, H.Z.; Simon, M.C.; Fredlund, E.; et al. Hypoxia-inducible factors 1 and 2 are important transcriptional effectors in primary macrophages experiencing hypoxia. Blood 2009, 114, 844–859. [Google Scholar] [CrossRef] [Green Version]
  130. Tafani, M.; Russo, A.; Di Vito, M.; Sale, P.; Pellegrini, L.; Schito, L.; Gentileschi, S.; Bracaglia, R.; Marandino, F.; Garaci, E.; et al. Up-regulation of pro-inflammatory genes as adaptation to hypoxia in MCF-7 cells and in human mammary invasive carcinoma microenvironment. Cancer Sci. 2010, 101, 1014–1023. [Google Scholar] [CrossRef] [PubMed]
  131. Raninga, P.V.; Di Trapani, G.; Vuckovic, S.; Tonissen, K.F. TrxR1 inhibition overcomes both hypoxia-induced and acquired bortezomib resistance in multiple myeloma through NF-кβ inhibition. Cell Cycle 2016, 15, 559–572. [Google Scholar] [CrossRef] [Green Version]
  132. Sakurai, A.; Yuasa, K.; Shoji, Y.; Himeno, S.; Tsujimoto, M.; Kunimoto, M.; Imura, N.; Hara, S. Overexpression of thioredoxin reductase 1 regulates NF-kappa B activation. J. Cell Physiol. 2004, 198, 22–30. [Google Scholar] [CrossRef]
  133. Liu, Z.B.; Shen, X. Thioredoxin reductase 1 upregulates MCP-1 release in human endothelial cells. Biochem. Biophys. Res. Commun. 2009, 386, 703–708. [Google Scholar] [CrossRef] [PubMed]
  134. Han, S.; Xu, W.; Wang, Z.; Qi, X.; Wang, Y.; Ni, Y.; Shen, H.; Hu, Q.; Han, W. Crosstalk between the HIF-1 and Toll-like receptor/nuclear factor-κB pathways in the oral squamous cell carcinoma microenvironment. Oncotarget 2016, 7, 37773–37789. [Google Scholar] [CrossRef] [PubMed]
  135. Ullah, K.; Rosendahl, A.H.; Izzi, V.; Bergmann, U.; Pihlajaniemi, T.; Mäki, J.M.; Myllyharju, J. Hypoxia-inducible factor prolyl-4-hydroxylase-1 is a convergent point in the reciprocal negative regulation of NF-κB and p53 signaling pathways. Sci. Rep. 2017, 7, 17220. [Google Scholar] [CrossRef] [Green Version]
  136. Scholz, C.C.; Cavadas, M.A.; Tambuwala, M.M.; Hams, E.; Rodríguez, J.; von Kriegsheim, A.; Cotter, P.; Bruning, U.; Fallon, P.G.; Cheong, A.; et al. Regulation of IL-1β-induced NF-κB by hydroxylases links key hypoxic and inflammatory signaling pathways. Proc. Natl. Acad. Sci. USA 2013, 110, 18490–18495. [Google Scholar] [CrossRef] [Green Version]
  137. Bandarra, D.; Biddlestone, J.; Mudie, S.; Müller, H.A.; Rocha, S. HIF-1α restricts NF-κB-dependent gene expression to control innate immunity signals. Dis. Model. Mech. 2015, 8, 169–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Müller-Edenborn, K.; Léger, K.; Glaus Garzon, J.F.; Oertli, C.; Mirsaidi, A.; Richards, P.J.; Rehrauer, H.; Spielmann, P.; Hoogewijs, D.; Borsig, L.; et al. Hypoxia attenuates the proinflammatory response in colon cancer cells by regulating IκB. Oncotarget 2015, 6, 20288–20301. [Google Scholar] [CrossRef] [Green Version]
  139. Mendonça, D.B.; Mendonça, G.; Aragão, F.J.; Cooper, L.F. NF-κB suppresses HIF-1α response by competing for P300 binding. Biochem. Biophys. Res. Commun. 2011, 404, 997–1003. [Google Scholar] [CrossRef]
  140. Mendonça, D.B.; Mendonça, G.; Cooper, L.F. Mammalian two-hybrid assays for studies of interaction of p300 with transcription factors. Methods Mol. Biol. 2013, 977, 323–338. [Google Scholar]
  141. Shin, D.H.; Li, S.H.; Yang, S.W.; Lee, B.L.; Lee, M.K.; Park, J.W. Inhibitor of nuclear factor-kappaB alpha derepresses hypoxia-inducible factor-1 during moderate hypoxia by sequestering factor inhibiting hypoxia-inducible factor from hypoxia-inducible factor 1alpha. FEBS J. 2009, 276, 3470–3480. [Google Scholar] [CrossRef]
  142. Xue, J.; Li, X.; Jiao, S.; Wei, Y.; Wu, G.; Fang, J. Prolyl hydroxylase-3 is down-regulated in colorectal cancer cells and inhibits IKKbeta independent of hydroxylase activity. Gastroenterology 2010, 138, 606–615. [Google Scholar] [CrossRef] [PubMed]
  143. D’Ignazio, L.; Shakir, D.; Batie, M.; Muller, H.A.; Rocha, S. HIF-1β Positively Regulates NF-κB Activity via Direct Control of TRAF6. Int. J. Mol. Sci. 2020, 21, 3000. [Google Scholar] [CrossRef] [PubMed]
  144. Dewhirst, M.W.; Braun, R.D.; Lanzen, J.L. Temporal changes in PO2 of R3230AC tumors in Fischer-344 rats. Int. J. Radiat. Oncol. Biol. Phys. 1998, 42, 723–726. [Google Scholar] [CrossRef]
  145. Brurberg, K.G.; Graff, B.A.; Olsen, D.R.; Rofstad, E.K. Tumor-line specific pO(2) fluctuations in human melanoma xenografts. Int. J. Radiat. Oncol. Biol. Phys. 2004, 58, 403–409. [Google Scholar] [CrossRef]
  146. Span, P.N.; Bussink, J. Biology of hypoxia. Semin. Nucl. Med. 2015, 45, 101–109. [Google Scholar] [CrossRef]
  147. Baluk, P.; Morikawa, S.; Haskell, A.; Mancuso, M.; McDonald, D.M. Abnormalities of basement membrane on blood vessels and endothelial sprouts in tumors. Am. J. Pathol. 2003, 163, 1801–1815. [Google Scholar] [CrossRef] [Green Version]
  148. Baluk, P.; Hashizume, H.; McDonald, D.M. Cellular abnormalities of blood vessels as targets in cancer. Curr. Opin. Genet. Dev. 2005, 15, 102–111. [Google Scholar] [CrossRef]
  149. Hashizume, H.; Baluk, P.; Morikawa, S.; McLean, J.W.; Thurston, G.; Roberge, S.; Jain, R.K.; McDonald, D.M. Openings between defective endothelial cells explain tumor vessel leakiness. Am. J. Pathol. 2000, 156, 1363–1380. [Google Scholar] [CrossRef] [Green Version]
  150. Lanzen, J.; Braun, R.D.; Klitzman, B.; Brizel, D.; Secomb, T.W.; Dewhirst, M.W. Direct demonstration of instabilities in oxygen concentrations within the extravascular compartment of an experimental tumor. Cancer Res. 2006, 66, 2219–2223. [Google Scholar] [CrossRef] [Green Version]
  151. Cárdenas-Navia, L.I.; Mace, D.; Richardson, R.A.; Wilson, D.F.; Shan, S.; Dewhirst, M.W. The pervasive presence of fluctuating oxygenation in tumors. Cancer Res. 2008, 68, 5812–5819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Baudelet, C.; Cron, G.O.; Ansiaux, R.; Crokart, N.; DeWever, J.; Feron, O.; Gallez, B. The role of vessel maturation and vessel functionality in spontaneous fluctuations of T2*-weighted GRE signal within tumors. NMR Biomed. 2006, 19, 69–76. [Google Scholar] [CrossRef]
  153. Panek, R.; Welsh, L.; Baker, L.C.J.; Schmidt, M.A.; Wong, K.H.; Riddell, A.M.; Koh, D.M.; Dunlop, A.; Mcquaid, D.; d’Arcy, J.A.; et al. Noninvasive Imaging of Cycling Hypoxia in Head and Neck Cancer Using Intrinsic Susceptibility MRI. Clin. Cancer Res. 2017, 23, 4233–4241. [Google Scholar] [CrossRef] [Green Version]
  154. Ellingsen, C.; Ovrebø, K.M.; Galappathi, K.; Mathiesen, B.; Rofstad, E.K. pO₂ fluctuation pattern and cycling hypoxia in human cervical carcinoma and melanoma xenografts. Int. J. Radiat. Oncol. Biol. Phys. 2012, 83, 1317–1323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Redler, G.; Epel, B.; Halpern, H.J. Principal component analysis enhances SNR for dynamic electron paramagnetic resonance oxygen imaging of cycling hypoxia in vivo. Magn. Reson. Med. 2014, 71, 440–450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Yasui, H.; Matsumoto, S.; Devasahayam, N.; Munasinghe, J.P.; Choudhuri, R.; Saito, K.; Subramanian, S.; Mitchell, J.B.; Krishna, M.C. Low-field magnetic resonance imaging to visualize chronic and cycling hypoxia in tumor-bearing mice. Cancer Res. 2010, 70, 6427–6436. [Google Scholar] [CrossRef] [Green Version]
  157. Zhang, J.; Zheng, L.; Cao, J.; Chen, B.; Jin, D. Inflammation induced by increased frequency of intermittent hypoxia is attenuated by tempol administration. Braz. J. Med. Biol. Res. 2015, 48, 1115–1121. [Google Scholar] [CrossRef] [Green Version]
  158. Yang, S.C.; Zhao, Y.; Zheng, Y.Y.; Li, W.Y.; Zhao, M.; Ji, E.S. Effects of intermittent hypoxia stimulation with different frequencies on HT22 cell viability and expression of Hif-1α and p-NF-κB. Acta Physiol Sinica 2021, 73, 26–34. [Google Scholar] [PubMed]
  159. Torres, M.; Martinez-Garcia, M.Á.; Campos-Rodriguez, F.; Gozal, D.; Montserrat, J.M.; Navajas, D.; Farré, R.; Almendros, I. Lung cancer aggressiveness in an intermittent hypoxia murine model of postmenopausal sleep apnea. Menopause 2020, 27, 706–713. [Google Scholar] [CrossRef]
  160. Hao, S.; Zhu, X.; Liu, Z.; Wu, X.; Li, S.; Jiang, P.; Jiang, L. Chronic intermittent hypoxia promoted lung cancer stem cell-like properties via enhancing Bach1 expression. Respir. Res. 2021, 22, 58. [Google Scholar] [CrossRef]
  161. Chen, W.L.; Wang, C.C.; Lin, Y.J.; Wu, C.P.; Hsieh, C.H. Cycling hypoxia induces chemoresistance through the activation of reactive oxygen species-mediated B-cell lymphoma extra-long pathway in glioblastoma multiforme. J. Transl. Med. 2015, 13, 389. [Google Scholar] [CrossRef] [Green Version]
  162. Cairns, R.A.; Kalliomaki, T.; Hill, R.P. Acute (cyclic) hypoxia enhances spontaneous metastasis of KHT murine tumors. Cancer Res. 2001, 61, 8903–8908. [Google Scholar]
  163. Miao, Z.F.; Zhao, T.T.; Wang, Z.N.; Xu, Y.Y.; Mao, X.Y.; Wu, J.H.; Liu, X.Y.; Xu, H.; You, Y.; Xu, H.M. Influence of different hypoxia models on metastatic potential of SGC-7901 gastric cancer cells. Tumour Biol. 2014, 35, 6801–6808. [Google Scholar] [CrossRef] [PubMed]
  164. Gu, X.; Zhang, J.; Shi, Y.; Shen, H.; Li, Y.; Chen, Y.; Liang, L. ESM1/HIF-1α pathway modulates chronic intermittent hypoxia-induced non-small-cell lung cancer proliferation, stemness and epithelial-mesenchymal transition. Oncol. Rep. 2021, 45, 1226–1234. [Google Scholar] [CrossRef] [PubMed]
  165. Delprat, V.; Tellier, C.; Demazy, C.; Raes, M.; Feron, O.; Michiels, C. Cycling hypoxia promotes a pro-inflammatory phenotype in macrophages via JNK/p65 signaling pathway. Sci. Rep. 2020, 10, 882. [Google Scholar] [CrossRef] [PubMed]
  166. Ke, X.; Chen, C.; Song, Y.; Cai, Q.; Li, J.; Tang, Y.; Han, X.; Qu, W.; Chen, A.; Wang, H.; et al. Hypoxia modifies the polarization of macrophages and their inflammatory microenvironment, and inhibits malignant behavior in cancer cells. Oncol. Lett. 2019, 18, 5871–5878. [Google Scholar] [CrossRef] [Green Version]
  167. Olbryt, M.; Habryka, A.; Student, S.; Jarząb, M.; Tyszkiewicz, T.; Lisowska, K.M. Global gene expression profiling in three tumor cell lines subjected to experimental cycling and chronic hypoxia. PLoS ONE 2014, 9, e105104. [Google Scholar]
  168. Tellier, C.; Desmet, D.; Petit, L.; Finet, L.; Graux, C.; Raes, M.; Feron, O.; Michiels, C. Cycling hypoxia induces a specific amplified inflammatory phenotype in endothelial cells and enhances tumor-promoting inflammation in vivo. Neoplasia 2015, 17, 66–78. [Google Scholar] [CrossRef] [Green Version]
  169. Gutsche, K.; Randi, E.B.; Blank, V.; Fink, D.; Wenger, R.H.; Leo, C.; Scholz, C.C. Intermittent hypoxia confers pro-metastatic gene expression selectively through NF-κB in inflammatory breast cancer cells. Free Radic. Biol. Med. 2016, 101, 129–142. [Google Scholar] [CrossRef] [Green Version]
  170. Li, D.; Wang, C.; Li, N.; Zhang, L. Propofol selectively inhibits nuclear factor-κB activity by suppressing p38 mitogen-activated protein kinase signaling in human EA.hy926 endothelial cells during intermittent hypoxia/reoxygenation. Mol. Med. Rep. 2014, 9, 1460–1466. [Google Scholar] [CrossRef] [Green Version]
  171. Hsieh, C.H.; Lee, C.H.; Liang, J.A.; Yu, C.Y.; Shyu, W.C. Cycling hypoxia increases U87 glioma cell radioresistance via ROS induced higher and long-term HIF-1 signal transduction activity. Oncol. Rep. 2010, 24, 1629–1636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Hsieh, C.H.; Shyu, W.C.; Chiang, C.Y.; Kuo, J.W.; Shen, W.C.; Liu, R.S. NADPH oxidase subunit 4-mediated reactive oxygen species contribute to cycling hypoxia-promoted tumor progression in glioblastoma multiforme. PLoS ONE 2011, 6, e23945. [Google Scholar] [CrossRef]
  173. Malec, V.; Gottschald, O.R.; Li, S.; Rose, F.; Seeger, W.; Hänze, J. HIF-1 alpha signaling is augmented during intermittent hypoxia by induction of the Nrf2 pathway in NOX1-expressing adenocarcinoma A549 cells. Free Radic. Biol. Med. 2010, 48, 1626–1635. [Google Scholar] [CrossRef] [PubMed]
  174. Li, L.; Ren, F.; Qi, C.; Xu, L.; Fang, Y.; Liang, M.; Feng, J.; Chen, B.; Ning, W.; Cao, J. Intermittent hypoxia promotes melanoma lung metastasis via oxidative stress and inflammation responses in a mouse model of obstructive sleep apnea. Respir. Res. 2018, 19, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Hsieh, C.H.; Chang, H.T.; Shen, W.C.; Shyu, W.C.; Liu, R.S. Imaging the impact of Nox4 in cycling hypoxia-mediated U87 glioblastoma invasion and infiltration. Mol. Imaging Biol. 2012, 14, 489–499. [Google Scholar] [CrossRef] [PubMed]
  176. Nanduri, J.; Vaddi, D.R.; Khan, S.A.; Wang, N.; Makerenko, V.; Prabhakar, N.R. Xanthine oxidase mediates hypoxia-inducible factor-2α degradation by intermittent hypoxia. PLoS ONE 2013, 8, e75838. [Google Scholar] [CrossRef]
  177. Nanduri, J.; Vaddi, D.R.; Khan, S.A.; Wang, N.; Makarenko, V.; Semenza, G.L.; Prabhakar, N.R. HIF-1α activation by intermittent hypoxia requires NADPH oxidase stimulation by xanthine oxidase. PLoS ONE 2015, 10, e0119762. [Google Scholar] [CrossRef] [Green Version]
  178. Yuan, G.; Adhikary, G.; McCormick, A.A.; Holcroft, J.J.; Kumar, G.K.; Prabhakar, N.R. Role of oxidative stress in intermittent hypoxia-induced immediate early gene activation in rat PC12 cells. J. Physiol. 2004, 557, 773–783. [Google Scholar] [CrossRef]
  179. Chen, X.; Li, X.; Zhang, W.; He, J.; Xu, B.; Lei, B.; Wang, Z.; Cates, C.; Rousselle, T.; Li, J. Activation of AMPK inhibits inflammatory response during hypoxia and reoxygenation through modulating JNK-mediated NF-κB pathway. Metabolism 2018, 83, 256–270. [Google Scholar] [CrossRef]
  180. Yuan, G.; Nanduri, J.; Khan, S.; Semenza, G.L.; Prabhakar, N.R. Induction of HIF-1alpha expression by intermittent hypoxia: Involvement of NADPH oxidase, Ca2+ signaling, prolyl hydroxylases, and mTOR. J. Cell Physiol. 2008, 217, 674–685. [Google Scholar] [CrossRef] [Green Version]
  181. Richard, D.E.; Berra, E.; Gothié, E.; Roux, D.; Pouysségur, J. p42/p44 mitogen-activated protein kinases phosphorylate hypoxia-inducible factor 1alpha (HIF-1alpha) and enhance the transcriptional activity of HIF-1. J. Biol. Chem. 1999, 274, 32631–32637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Liu, C.; Shi, Y.; Han, Z.; Pan, Y.; Liu, N.; Han, S.; Chen, Y.; Lan, M.; Qiao, T.; Fan, D. Suppression of the dual-specificity phosphatase MKP-1 enhances HIF-1 trans-activation and increases expression of EPO. Biochem. Biophys. Res. Commun. 2003, 312, 780–786. [Google Scholar] [CrossRef] [PubMed]
  183. Koga, S.; Kojima, S.; Kishimoto, T.; Kuwabara, S.; Yamaguchi, A. Over-expression of map kinase phosphatase-1 (MKP-1) suppresses neuronal death through regulating JNK signaling in hypoxia/re-oxygenation. Brain Res. 2012, 1436, 137–146. [Google Scholar] [CrossRef] [PubMed]
  184. Hoffmann, M.S.; Singh, P.; Wolk, R.; Narkiewicz, K.; Somers, V.K. Obstructive sleep apnea and intermittent hypoxia increase expression of dual specificity phosphatase 1. Atherosclerosis 2013, 231, 378–383. [Google Scholar] [CrossRef] [Green Version]
  185. Toffoli, S.; Feron, O.; Raes, M.; Michiels, C. Intermittent hypoxia changes HIF-1alpha phosphorylation pattern in endothelial cells: Unravelling of a new PKA-dependent regulation of HIF-1alpha. Biochim. Biophys. Acta 2007, 1773, 1558–1571. [Google Scholar] [CrossRef] [Green Version]
  186. Bullen, J.W.; Tchernyshyov, I.; Holewinski, R.J.; DeVine, L.; Wu, F.; Venkatraman, V.; Kass, D.L.; Cole, R.N.; Van Eyk, J.; Semenza, G.L. Protein kinase A-dependent phosphorylation stimulates the transcriptional activity of hypoxia-inducible factor 1. Sci. Signal. 2016, 9, ra56. [Google Scholar] [CrossRef] [Green Version]
  187. Zhang, Y.L.; Tavakoli, H.; Chachisvilis, M. Apparent PKA activity responds to intermittent hypoxia in bone cells: A redox pathway? Am. J. Physiol. Heart Circ. Physiol. 2010, 299, H225–H235. [Google Scholar] [CrossRef] [Green Version]
  188. Naranjo-Suarez, S.; Carlson, B.A.; Tobe, R.; Yoo, M.H.; Tsuji, P.A.; Gladyshev, V.N.; Hatfield, D.L. Regulation of HIF-1α activity by overexpression of thioredoxin is independent of thioredoxin reductase status. Mol. Cells 2013, 36, 151–157. [Google Scholar] [CrossRef] [Green Version]
  189. Zhao, L.; Li, W.; Zhou, Y.; Zhang, Y.; Huang, S.; Xu, X.; Li, Z.; Guo, Q. The overexpression and nuclear translocation of Trx-1 during hypoxia confers on HepG2 cells resistance to DDP, and GL-V9 reverses the resistance by suppressing the Trx-1/Ref-1 axis. Free Radic. Biol. Med. 2015, 82, 29–41. [Google Scholar] [CrossRef]
  190. Wang, N.; Peng, Y.J.; Su, X.; Prabhakar, N.R.; Nanduri, J. Histone Deacetylase 5 Is an Early Epigenetic Regulator of Intermittent Hypoxia Induced Sympathetic Nerve Activation and Blood Pressure. Front. Physiol. 2021, 12, 688322. [Google Scholar] [CrossRef]
  191. Quintero, M.; Gonzalez-Martin, M.D.C.; Vega-Agapito, V.; Gonzalez, C.; Obeso, A.; Farré, R.; Agapito, T.; Yubero, S. The effects of intermittent hypoxia on redox status, NF-κB activation, and plasma lipid levels are dependent on the lowest oxygen saturation. Free Radic. Biol. Med. 2013, 65, 1143–1154. [Google Scholar] [CrossRef]
  192. Kunz, M.; Bloss, G.; Gillitzer, R.; Gross, G.; Goebeler, M.; Rapp, U.R.; Ludwig, S. Hypoxia/reoxygenation induction of monocyte chemoattractant protein-1 in melanoma cells: Involvement of nuclear factor-kappaB, stimulatory protein-1 transcription factors and mitogen-activated protein kinase pathways. Biochem. J. 2002, 366, 299–306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Ryan, S.; McNicholas, W.T.; Taylor, C.T. A critical role for p38 map kinase in NF-kappaB signaling during intermittent hypoxia/reoxygenation. Biochem. Biophys. Res. Commun. 2007, 355, 728–733. [Google Scholar] [CrossRef] [PubMed]
  194. Lee, M.Y.; Wang, Y.; Mak, J.C.; Ip, M.S. Intermittent hypoxia induces NF-κB-dependent endothelial activation via adipocyte-derived mediators. Am. J. Physiol. Cell Physiol. 2016, 310, C446–C455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Zhang, Y.; Luo, Y.; Wang, Y.; Liu, H.; Yang, Y.; Wang, Q. Effect of deubiquitinase USP8 on hypoxia/reoxygenation-induced inflammation by deubiquitination of TAK1 in renal tubular epithelial cells. Int. J. Mol. Med. 2018, 42, 3467–3476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Song, D.; Fang, G.; Mao, S.Z.; Ye, X.; Liu, G.; Miller, E.J.; Greenberg, H.; Liu, S.F. Selective inhibition of endothelial NF-κB signaling attenuates chronic intermittent hypoxia-induced atherosclerosis in mice. Atherosclerosis 2018, 270, 68–75. [Google Scholar] [CrossRef]
  197. Daneau, G.; Boidot, R.; Martinive, P.; Feron, O. Identification of cyclooxygenase-2 as a major actor of the transcriptomic adaptation of endothelial and tumor cells to cyclic hypoxia: Effect on angiogenesis and metastases. Clin. Cancer Res. 2010, 16, 410–419. [Google Scholar] [CrossRef] [Green Version]
  198. Naidu, B.V.; Krishnadasan, B.; Byrne, K.; Farr, A.L.; Rosengart, M.; Verrier, E.D.; Mulligan, M.S. Regulation of chemokine expression by cyclosporine A in alveolar macrophages exposed to hypoxia and reoxygenation. Ann. Thorac. Surg. 2002, 74, 899–905. [Google Scholar] [CrossRef]
  199. Chuang, L.P.; Chen, N.H.; Lin, Y.; Ko, W.S.; Pang, J.H. Increased MCP-1 gene expression in monocytes of severe OSA patients and under intermittent hypoxia. Sleep Breath 2016, 20, 425–433. [Google Scholar] [CrossRef]
  200. Higashihara, H.; Kokura, S.; Imamoto, E.; Ueda, M.; Naito, Y.; Yoshida, N.; Yoshikawa, T. Hypoxia-reoxygenation enhances interleukin-8 production from U937 human monocytic cells. Redox Rep. 2004, 9, 365–369. [Google Scholar] [CrossRef]
  201. Dyugovskaya, L.; Polyakov, A.; Ginsberg, D.; Lavie, P.; Lavie, L. Molecular pathways of spontaneous and TNF-{alpha}-mediated neutrophil apoptosis under intermittent hypoxia. Am. J. Respir. Cell Mol. Biol 2011, 45, 154–162. [Google Scholar] [CrossRef]
  202. Dong, G.; Lin, X.H.; Liu, H.H.; Gao, D.M.; Cui, J.F.; Ren, Z.G.; Chen, R.X. Intermittent hypoxia alleviates increased VEGF and pro-angiogenic potential in liver cancer cells. Oncol. Lett. 2019, 18, 1831–1839. [Google Scholar] [CrossRef] [PubMed]
  203. Rofstad, E.K.; Gaustad, J.V.; Egeland, T.A.; 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] [PubMed]
  204. Itatani, Y.; Kawada, K.; Yamamoto, T.; Sakai, Y. Resistance to Anti-Angiogenic Therapy in Cancer-Alterations to Anti-VEGF Pathway. Int. J. Mol. Sci. 2018, 19, 1232. [Google Scholar] [CrossRef] [Green Version]
  205. Melincovici, C.S.; Boşca, A.B.; Şuşman, S.; Mărginean, M.; Mihu, C.; Istrate, M.; Moldovan, I.M.; Roman, A.L.; Mihu, C.M. Vascular endothelial growth factor (VEGF)—Key factor in normal and pathological angiogenesis. Rom. J. Morphol. Embryol. 2018, 59, 455–467. [Google Scholar] [PubMed]
  206. Kuroda, T.; Kitadai, Y.; Tanaka, S.; Yang, X.; Mukaida, N.; Yoshihara, M.; Chayama, K. Monocyte chemoattractant protein-1 transfection induces angiogenesis and tumorigenesis of gastric carcinoma in nude mice via macrophage recruitment. Clin. Cancer Res. 2005, 11, 7629–7636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Varney, M.L.; Olsen, K.J.; Mosley, R.L.; Singh, R.K. Paracrine regulation of vascular endothelial growth factor—A expression during macrophage-melanoma cell interaction: Role of monocyte chemotactic protein-1 and macrophage colony-stimulating factor. J. Interferon. Cytokine Res. 2005, 25, 674–683. [Google Scholar] [CrossRef] [PubMed]
  208. Wang, R.; Zhang, J.; Chen, S.; Lu, M.; Luo, X.; Yao, S.; Liu, S.; Qin, Y.; Chen, H. Tumor-associated macrophages provide a suitable microenvironment for non-small lung cancer invasion and progression. Lung Cancer 2011, 74, 188–196. [Google Scholar] [CrossRef]
  209. Salcedo, R.; Ponce, M.L.; Young, H.A.; Wasserman, K.; Ward, J.M.; Kleinman, H.K.; Oppenheim, J.J.; Murphy, W.J. Human endothelial cells express CCR2 and respond to MCP-1: Direct role of MCP-1 in angiogenesis and tumor progression. Blood 2000, 96, 34–40. [Google Scholar] [CrossRef]
  210. Loetscher, P.; Seitz, M.; Clark-Lewis, I.; Baggiolini, M.; Moser, B. Both interleukin-8 receptors independently mediate chemotaxis. Jurkat cells transfected with IL-8R1 or IL-8R2 migrate in response to IL-8, GRO alpha and NAP-2. FEBS Lett. 1994, 341, 187–192. [Google Scholar] [CrossRef] [Green Version]
  211. Hughes, C.E.; Nibbs, R.J.B. A guide to chemokines and their receptors. FEBS J. 2018, 285, 2944–2971. [Google Scholar] [CrossRef]
  212. Keane, M.P.; Belperio, J.A.; Xue, Y.Y.; Burdick, M.D.; Strieter, R.M. Depletion of CXCR2 inhibits tumor growth and angiogenesis in a murine model of lung cancer. J. Immunol. 2004, 172, 2853–2860. [Google Scholar] [CrossRef] [Green Version]
  213. Strieter, R.M.; Burdick, M.D.; Mestas, J.; Gomperts, B.; Keane, M.P.; Belperio, J.A. Cancer CXC chemokine networks and tumour angiogenesis. Eur. J. Cancer 2006, 42, 768–778. [Google Scholar] [CrossRef]
  214. Liu, L.; Sun, H.; Wu, S.; Tan, H.; Sun, Y.; Liu, X.; Si, S.; Xu, L.; Huang, J.; Zhou, W.; et al. IL-17A promotes CXCR2-dependent angiogenesis in a mouse model of liver cancer. Mol. Med. Rep. 2019, 20, 1065–1074. [Google Scholar] [CrossRef] [Green Version]
  215. Haqqani, A.S.; Sandhu, J.K.; Birnboim, H.C. Expression of interleukin-8 promotes neutrophil infiltration and genetic instability in mutatect tumors. Neoplasia 2000, 2, 561–568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Yao, C.; Lin, Y.; Chua, M.S.; Ye, C.S.; Bi, J.; Li, W.; Zhu, Y.F.; Wang, S.M. Interleukin-8 modulates growth and invasiveness of estrogen receptor-negative breast cancer cells. Int. J. Cancer 2007, 121, 1949–1957. [Google Scholar] [CrossRef]
  217. Jablonska, J.; Wu, C.F.; Andzinski, L.; Leschner, S.; Weiss, S. CXCR2-mediated tumor-associated neutrophil recruitment is regulated by IFN-β. Int. J. Cancer 2014, 134, 1346–1358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Yuan, M.; Zhu, H.; Xu, J.; Zheng, Y.; Cao, X.; Liu, Q. Tumor-Derived CXCL1 Promotes Lung Cancer Growth via Recruitment of Tumor-Associated Neutrophils. J. Immunol. Res. 2016, 2016, 6530410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Bekes, E.M.; Schweighofer, B.; Kupriyanova, T.A.; Zajac, E.; Ardi, V.C.; Quigley, J.P.; Deryugina, E.I. Tumor-recruited neutrophils and neutrophil TIMP-free MMP-9 regulate coordinately the levels of tumor angiogenesis and efficiency of malignant cell intravasation. Am. J. Pathol. 2011, 179, 1455–1470. [Google Scholar] [CrossRef] [PubMed]
  220. Deryugina, E.I.; Zajac, E.; Juncker-Jensen, A.; Kupriyanova, T.A.; Welter, L.; Quigley, J.P. Tissue-infiltrating neutrophils constitute the major in vivo source of angiogenesis-inducing MMP-9 in the tumor microenvironment. Neoplasia 2014, 16, 771–788. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Hawinkels, L.J.; Zuidwijk, K.; Verspaget, H.W.; de Jonge-Muller, E.S.; van Duijn, W.; Ferreira, V.; Fontijn, R.D.; David, G.; Hommes, D.W.; Lamers, C.B.; et al. VEGF release by MMP-9 mediated heparan sulphate cleavage induces colorectal cancer angiogenesis. Eur. J. Cancer 2008, 44, 1904–1913. [Google Scholar] [CrossRef]
  222. Tsujii, M.; Kawano, S.; Tsuji, S.; Sawaoka, H.; Hori, M.; DuBois, R.N. Cyclooxygenase regulates angiogenesis induced by colon cancer cells. Cell 1998, 93, 705–716. [Google Scholar] [CrossRef] [Green Version]
  223. Liu, H.; Xiao, J.; Yang, Y.; Liu, Y.; Ma, R.; Li, Y.; Deng, F.; Zhang, Y. COX-2 expression is correlated with VEGF-C, lymphangiogenesis and lymph node metastasis in human cervical cancer. Microvasc. Res. 2011, 82, 131–140. [Google Scholar] [CrossRef]
  224. Xin, X.; Majumder, M.; Girish, G.V.; Mohindra, V.; Maruyama, T.; Lala, P.K. Targeting COX-2 and EP4 to control tumor growth, angiogenesis, lymphangiogenesis and metastasis to the lungs and lymph nodes in a breast cancer model. Lab. Investig. 2012, 92, 1115–1128. [Google Scholar] [CrossRef] [Green Version]
  225. Zhao, L.; Wu, Y.; Xu, Z.; Wang, H.; Zhao, Z.; Li, Y.; Yang, P.; Wei, X. Involvement of COX-2/PGE2 signalling in hypoxia-induced angiogenic response in endothelial cells. J. Cell Mol. Med. 2012, 16, 1840–1855. [Google Scholar] [CrossRef] [PubMed]
  226. Xie, C.; Xu, X.; Wang, X.; Wei, S.; Shao, L.; Chen, J.; Cai, J.; Jia, L. Cyclooxygenase-2 induces angiogenesis in pancreatic cancer mediated by prostaglandin E2. Oncol. Lett. 2018, 16, 940–948. [Google Scholar] [CrossRef] [Green Version]
  227. Salcedo, R.; Zhang, X.; Young, H.A.; Michael, N.; Wasserman, K.; Ma, W.H.; Martins-Green, M.; Murphy, W.J.; Oppenheim, J.J. Angiogenic effects of prostaglandin E2 are mediated by up-regulation of CXCR4 on human microvascular endothelial cells. Blood 2003, 102, 1966–1977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Grossman, J.G.; Nywening, T.M.; Belt, B.A.; Panni, R.Z.; Krasnick, B.A.; DeNardo, D.G.; Hawkins, W.G.; Goedegebuure, S.P.; Linehan, D.C.; Fields, R.C. Recruitment of CCR2+ tumor associated macrophage to sites of liver metastasis confers a poor prognosis in human colorectal cancer. Oncoimmunology 2018, 7, e1470729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Li, F.; Kitajima, S.; Kohno, S.; Yoshida, A.; Tange, S.; Sasaki, S.; Okada, N.; Nishimoto, Y.; Muranaka, H.; Nagatani, N.; et al. Retinoblastoma Inactivation Induces a Protumoral Microenvironment via Enhanced CCL2 Secretion. Cancer Res. 2019, 79, 3903–3915. [Google Scholar] [CrossRef] [Green Version]
  230. Muthuswamy, R.; Urban, J.; Lee, J.J.; Reinhart, T.A.; Bartlett, D.; Kalinski, P. Ability of mature dendritic cells to interact with regulatory T cells is imprinted during maturation. Cancer Res. 2008, 68, 5972–5978. [Google Scholar] [CrossRef] [Green Version]
  231. Baratelli, F.; Lee, J.M.; Hazra, S.; Lin, Y.; Walser, T.C.; Schaue, D.; Pak, P.S.; Elashoff, D.; Reckamp, K.; Zhang, L.; et al. PGE(2) contributes to TGF-beta induced T regulatory cell function in human non-small cell lung cancer. Am. J. Transl. Res. 2010, 2, 356–367. [Google Scholar]
  232. Yuan, X.L.; Chen, L.; Li, M.X.; Dong, P.; Xue, J.; Wang, J.; Zhang, T.T.; Wang, X.A.; Zhang, F.M.; Ge, H.L.; et al. Elevated expression of Foxp3 in tumor-infiltrating Treg cells suppresses T-cell proliferation and contributes to gastric cancer progression in a COX-2-dependent manner. Clin. Immunol. 2010, 134, 277–288. [Google Scholar] [CrossRef]
  233. Prima, V.; Kaliberova, L.N.; Kaliberov, S.; Curiel, D.T.; Kusmartsev, S. COX2/mPGES1/PGE2 pathway regulates PD-L1 expression in tumor-associated macrophages and myeloid-derived suppressor cells. Proc. Natl. Acad. Sci. USA 2017, 114, 1117–1122. [Google Scholar] [CrossRef] [Green Version]
  234. Park, A.; Lee, Y.; Kim, M.S.; Kang, Y.J.; Park, Y.J.; Jung, H.; Kim, T.D.; Lee, H.G.; Choi, I.; Yoon, S.R. Prostaglandin E2 Secreted by Thyroid Cancer Cells Contributes to Immune Escape Through the Suppression of Natural Killer (NK) Cell Cytotoxicity and NK Cell Differentiation. Front. Immunol. 2018, 9, 1859. [Google Scholar] [CrossRef] [PubMed]
  235. Rothwell, P.M.; Fowkes, F.G.; Belch, J.F.; Ogawa, H.; Warlow, C.P.; Meade, T.W. Effect of daily aspirin on long-term risk of death due to cancer: Analysis of individual patient data from randomised trials. Lancet 2011, 377, 31–41. [Google Scholar] [CrossRef]
  236. de Pedro, M.; Baeza, S.; Escudero, M.T.; Dierssen-Sotos, T.; Gómez-Acebo, I.; Pollán, M.; Llorca, J. Effect of COX-2 inhibitors and other non-steroidal inflammatory drugs on breast cancer risk: A meta-analysis. Breast Cancer Res. Treat. 2015, 149, 525–536. [Google Scholar] [CrossRef] [PubMed]
  237. Liu, Y.; Chen, J.Q.; Xie, L.; Wang, J.; Li, T.; He, Y.; Gao, Y.; Qin, X.; Li, S. Effect of aspirin and other non-steroidal anti-inflammatory drugs on prostate cancer incidence and mortality: A systematic review and meta-analysis. BMC Med. 2014, 12, 55. [Google Scholar] [CrossRef] [PubMed]
  238. Dai, P.; Li, J.; Ma, X.P.; Huang, J.; Meng, J.J.; Gong, P. Efficacy and safety of COX-2 inhibitors for advanced non-small-cell lung cancer with chemotherapy: A meta-analysis. Onco. Targets Ther. 2018, 11, 721–730. [Google Scholar] [CrossRef] [Green Version]
  239. Yi, L.; Zhang, W.; Zhang, H.; Shen, J.; Zou, J.; Luo, P.; Zhang, J. Systematic review and meta-analysis of the benefit of celecoxib in treating advanced non-small-cell lung cancer. Drug Des. Dev. Ther. 2018, 12, 2455–2466. [Google Scholar] [CrossRef] [Green Version]
  240. Xu, Y.Q.; Long, X.; Han, M.; Huang, M.Q.; Lu, J.F.; Sun, X.D.; Han, W. Clinical benefit of COX-2 inhibitors in the adjuvant chemotherapy of advanced non-small cell lung cancer: A systematic review and meta-analysis. World J. Clin. Cases 2021, 9, 581–601. [Google Scholar] [CrossRef]
  241. Popivanova, B.K.; Kostadinova, F.I.; Furuichi, K.; Shamekh, M.M.; Kondo, T.; Wada, T.; Egashira, K.; Mukaida, N. Blockade of a chemokine, CCL2, reduces chronic colitis-associated carcinogenesis in mice. Cancer Res. 2009, 69, 7884–7892. [Google Scholar] [CrossRef] [Green Version]
  242. An, J.; Xue, Y.; Long, M.; Zhang, G.; Zhang, J.; Su, H. Targeting CCR2 with its antagonist suppresses viability, motility and invasion by downregulating MMP-9 expression in non-small cell lung cancer cells. Oncotarget 2017, 8, 39230–39240. [Google Scholar] [CrossRef] [Green Version]
  243. Tu, M.M.; Abdel-Hafiz, H.A.; Jones, R.T.; Jean, A.; Hoff, K.J.; Duex, J.E.; Chauca-Diaz, A.; Costello, J.C.; Dancik, G.M.; Tamburini, B.A.J.; et al. Inhibition of the CCL2 receptor, CCR2, enhances tumor response to immune checkpoint therapy. Commun. Biol. 2020, 3, 720. [Google Scholar] [CrossRef]
  244. Loberg, R.D.; Ying, C.; Craig, M.; Day, L.L.; Sargent, E.; Neeley, C.; Wojno, K.; Snyder, L.A.; Yan, L.; Pienta, K.J. Targeting CCL2 with systemic delivery of neutralizing antibodies induces prostate cancer tumor regression in vivo. Cancer Res. 2007, 67, 9417–9424. [Google Scholar] [CrossRef] [Green Version]
  245. Rozel, S.; Galbán, C.J.; Nicolay, K.; Lee, K.C.; Sud, S.; Neeley, C.; Snyder, L.A.; Chenevert, T.L.; Rehemtulla, A.; Ross, B.D.; et al. Synergy between anti-CCL2 and docetaxel as determined by DW-MRI in a metastatic bone cancer model. J. Cell Biochem. 2009, 107, 58–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Sandhu, S.K.; Papadopoulos, K.; Fong, P.C.; Patnaik, A.; Messiou, C.; Olmos, D.; Wang, G.; Tromp, B.J.; Puchalski, T.A.; Balkwill, F.; et al. A first-in-human, first-in-class, phase I study of carlumab (CNTO 888), a human monoclonal antibody against CC-chemokine ligand 2 in patients with solid tumors. Cancer Chemother. Pharm. 2013, 71, 1041–1050. [Google Scholar] [CrossRef] [PubMed]
  247. Teng, K.Y.; Han, J.; Zhang, X.; Hsu, S.H.; He, S.; Wani, N.A.; Barajas, J.M.; Snyder, L.A.; Frankel, W.L.; Caligiuri, M.A.; et al. Blocking the CCL2-CCR2 Axis Using CCL2-Neutralizing Antibody Is an Effective Therapy for Hepatocellular Cancer in a Mouse Model. Mol. Cancer 2017, 16, 312–322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Miyake, M.; Furuya, H.; Onishi, S.; Hokutan, K.; Anai, S.; Chan, O.; Shi, S.; Fujimoto, K.; Goodison, S.; Cai, W.; et al. Monoclonal Antibody against CXCL1 (HL2401) as a Novel Agent in Suppressing IL6 Expression and Tumoral Growth. Theranostics 2019, 9, 853–867. [Google Scholar] [CrossRef] [PubMed]
  249. Mian, B.M.; Dinney, C.P.; Bermejo, C.E.; Sweeney, P.; Tellez, C.; Yang, X.D.; Gudas, J.M.; McConkey, D.J.; Bar-Eli, M. Fully human anti-interleukin 8 antibody inhibits tumor growth in orthotopic bladder cancer xenografts via down-regulation of matrix metalloproteases and nuclear factor-kappaB. Clin. Cancer Res. 2003, 9, 3167–3175. [Google Scholar] [PubMed]
  250. Wu, S.; Shang, H.; Cui, L.; Zhang, Z.; Zhang, Y.; Li, Y.; Wu, J.; Li, R.K.; Xie, J. Targeted blockade of interleukin-8 abrogates its promotion of cervical cancer growth and metastasis. Mol. Cell Biochem. 2013, 375, 69–79. [Google Scholar] [CrossRef] [PubMed]
  251. Dominguez, C.; McCampbell, K.K.; David, J.M.; Palena, C. Neutralization of IL-8 decreases tumor PMN-MDSCs and reduces mesenchymalization of claudin-low triple-negative breast cancer. JCI Insight 2017, 2, e94296. [Google Scholar] [CrossRef] [PubMed]
  252. Bilusic, M.; Heery, C.R.; Collins, J.M.; Donahue, R.N.; Palena, C.; Madan, R.A.; Karzai, F.; Marté, J.L.; Strauss, J.; Gatti-Mays, M.E.; et al. Phase I trial of HuMax-IL8 (BMS-986253), an anti-IL-8 monoclonal antibody, in patients with metastatic or unresectable solid tumors. J. Immunother. Cancer 2019, 7, 240. [Google Scholar] [CrossRef] [PubMed]
  253. Du, M.; Qiu, Q.; Gruslin, A.; Gordon, J.; He, M.; Chan, C.C.; Li, D.; Tsang, B.K. SB225002 promotes mitotic catastrophe in chemo-sensitive and -resistant ovarian cancer cells independent of p53 status in vitro. PLoS ONE 2013, 8, e54572. [Google Scholar] [CrossRef] [PubMed]
  254. Devapatla, B.; Sharma, A.; Woo, S. CXCR2 Inhibition Combined with Sorafenib Improved Antitumor and Antiangiogenic Response in Preclinical Models of Ovarian Cancer. PLoS ONE 2015, 10, e0139237. [Google Scholar] [CrossRef] [Green Version]
  255. Xu, M.; Jiang, H.; Wang, H.; Liu, J.; Liu, B.; Guo, Z. SB225002 inhibits prostate cancer invasion and attenuates the expression of BSP, OPN and MMP-2. Oncol. Rep. 2018, 40, 726–736. [Google Scholar] [CrossRef]
  256. Ruiz de Porras, V.; Wang, X.C.; Palomero, L.; Marin-Aguilera, M.; Solé-Blanch, C.; Indacochea, A.; Jimenez, N.; Bystrup, S.; Bakht, M.; Conteduca, V.; et al. Taxane-induced Attenuation of the CXCR2/BCL-2 Axis Sensitizes Prostate Cancer to Platinum-based Treatment. Eur. Urol. 2021, 79, 722–733. [Google Scholar] [CrossRef]
  257. Li, L.; Khan, M.N.; Li, Q.; Chen, X.; Wei, J.; Wang, B.; Cheng, J.W.; Gordon, J.R.; Li, F. G31P, CXCR1/2 inhibitor, with cisplatin inhibits the growth of mice hepatocellular carcinoma and mitigates high-dose cisplatin-induced nephrotoxicity. Oncol. Rep. 2015, 33, 751–757. [Google Scholar] [CrossRef]
  258. Wang, J.; Hu, W.; Wang, K.; Yu, J.; Luo, B.; Luo, G.; Wang, W.; Wang, H.; Li, J.; Wen, J. Repertaxin, an inhibitor of the chemokine receptors CXCR1 and CXCR2, inhibits malignant behavior of human gastric cancer MKN45 cells in vitro and in vivo and enhances efficacy of 5-fluorouracil. Int. J. Oncol. 2016, 48, 1341–1352. [Google Scholar] [CrossRef]
  259. Kemp, D.M.; Pidich, A.; Larijani, M.; Jonas, R.; Lash, E.; Sato, T.; Terai, M.; De Pizzol, M.; Allegretti, M.; Igoucheva, O.; et al. Ladarixin, a dual CXCR1/2 inhibitor, attenuates experimental melanomas harboring different molecular defects by affecting malignant cells and tumor microenvironment. Oncotarget 2017, 8, 14428–14442. [Google Scholar] [CrossRef] [Green Version]
  260. Goldstein, L.J.; Perez, R.P.; Yardley, D.; Han, L.K.; Reuben, J.M.; Gao, H.; McCanna, S.; Butler, B.; Ruffini, P.A.; Liu, Y.; et al. A window-of-opportunity trial of the CXCR1/2 inhibitor reparixin in operable HER-2-negative breast cancer. Breast Cancer Res. 2020, 22, 4. [Google Scholar] [CrossRef]
  261. Greene, S.; Robbins, Y.; Mydlarz, W.K.; Huynh, A.P.; Schmitt, N.C.; Friedman, J.; Horn, L.A.; Palena, C.; Schlom, J.; Maeda, D.Y.; et al. Inhibition of MDSC Trafficking with SX-682, a CXCR1/2 Inhibitor, Enhances NK-Cell Immunotherapy in Head and Neck Cancer Models. Clin. Cancer Res. 2020, 26, 1420–1431. [Google Scholar] [CrossRef] [Green Version]
  262. Erstad, D.J.; Cusack, J.C., Jr. Targeting the NF-κB pathway in cancer therapy. Surg. Oncol. Clin. 2013, 22, 705–746. [Google Scholar] [CrossRef]
  263. Garcia, J.; Hurwitz, H.I.; Sandler, A.B.; Miles, D.; Coleman, R.L.; Deurloo, R.; Chinot, O.L. Bevacizumab (Avastin®) in cancer treatment: A review of 15 years of clinical experience and future outlook. Cancer Treat. Rev. 2020, 86, 102017. [Google Scholar] [CrossRef] [PubMed]
  264. Schultheis, A.M.; Lurje, G.; Rhodes, K.E.; Zhang, W.; Yang, D.; Garcia, A.A.; Morgan, R.; Gandara, D.; Scudder, S.; Oza, A.; et al. Polymorphisms and clinical outcome in recurrent ovarian cancer treated with cyclophosphamide and bevacizumab. Clin. Cancer Res. 2008, 14, 7554–7563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Feng, H.; Liu, K.; Shen, X.; Liang, J.; Wang, C.; Qiu, W.; Cheng, X.; Zhao, R. Targeting tumor cell-derived CCL2 as a strategy to overcome Bevacizumab resistance in ETV5+ colorectal cancer. Cell Death Dis. 2020, 11, 916. [Google Scholar] [CrossRef] [PubMed]
  266. Abd-Rabou, A.A.; Ahmed, H.H. Bevacizumab and CCR2 Inhibitor Nanoparticles Induce Cytotoxicity-Mediated Apoptosis in Doxorubicin-Treated Hepatic and Non-Small Lung Cancer Cells. Asian Pac. J. Cancer Prev. 2019, 20, 2225–2238. [Google Scholar] [CrossRef]
  267. Xu, L.; Croix, B.S. Improving VEGF-targeted therapies through inhibition of COX-2/PGE2 signaling. Mol. Cell Oncol. 2014, 1, e969154. [Google Scholar] [CrossRef] [Green Version]
  268. Carbone, C.; Tamburrino, A.; Piro, G.; Boschi, F.; Cataldo, I.; Zanotto, M.; Mina, M.M.; Zanini, S.; Sbarbati, A.; Scarpa, A.; et al. Combined inhibition of IL1, CXCR1/2, and TGFβ signaling pathways modulates in-vivo resistance to anti-VEGF treatment. Anticancer Drugs 2016, 27, 29–40. [Google Scholar] [CrossRef]
  269. Cusack, J.C., Jr.; Liu, R.; Xia, L.; Chao, T.H.; Pien, C.; Niu, W.; Palombella, V.J.; Neuteboom, S.T.; Palladino, M.A. NPI-0052 enhances tumoricidal response to conventional cancer therapy in a colon cancer model. Clin. Cancer Res. 2006, 12, 6758–6764. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The effects of ROS and NO on the activation of HIF-1 in chronic hypoxia. Chronic hypoxia there is associated with an increase in the level of ROS which inactivate FIH and PHD. This increases the activation of HIF-1. ROS are also involved in the activation of NF-κB, a transcription factor important in the full activation of HIF-1. HIF-1 activation can also be induced by NO, especially at sites of inflammatory reactions. NO causes the S-nitrosylation of HIF-1α, which increases the stability of this protein. Another post-translational modification of HIF-1α induced by NO is phosphorylation associated with the inactivation of DUSP1. NO can also bind to the iron atom in PHDs and thus inactivate these enzymes. However, in combination with ROS, NO can restore activity of PHDs in chronic hypoxia. It can also increase calcium ion levels in the cytoplasm which activates calpain—a protease that degrades HIF-1α independently of the 26S proteasome.
Figure 1. The effects of ROS and NO on the activation of HIF-1 in chronic hypoxia. Chronic hypoxia there is associated with an increase in the level of ROS which inactivate FIH and PHD. This increases the activation of HIF-1. ROS are also involved in the activation of NF-κB, a transcription factor important in the full activation of HIF-1. HIF-1 activation can also be induced by NO, especially at sites of inflammatory reactions. NO causes the S-nitrosylation of HIF-1α, which increases the stability of this protein. Another post-translational modification of HIF-1α induced by NO is phosphorylation associated with the inactivation of DUSP1. NO can also bind to the iron atom in PHDs and thus inactivate these enzymes. However, in combination with ROS, NO can restore activity of PHDs in chronic hypoxia. It can also increase calcium ion levels in the cytoplasm which activates calpain—a protease that degrades HIF-1α independently of the 26S proteasome.
Ijms 22 10701 g001
Figure 2. The importance of MAPK cascades for the HIF-1 activation pathway. During chronic hypoxia, MAPK cascades, in particular ERK MAPK and p38 MAPK, are activated by ROS and increased calcium ions. These kinases cause phosphorylation of HIF-1α and consequently increase the stability and transcriptional activity of HIF-1. p38 MAPK can also cause the activation of SIAH2, which results in the ubiquitination and degradation of PHD3. Important in this model of HIF-1 activation are also phosphatases, in particular DUSP1 and DUSP2—enzymes that catalyze a reaction reverse to ERK MAPK and p38 MAPK. In chronic hypoxia, there is a decrease in DUSP2 expression but an increase in DUSP1, which is a mechanism for regulating HIF-1 activation.
Figure 2. The importance of MAPK cascades for the HIF-1 activation pathway. During chronic hypoxia, MAPK cascades, in particular ERK MAPK and p38 MAPK, are activated by ROS and increased calcium ions. These kinases cause phosphorylation of HIF-1α and consequently increase the stability and transcriptional activity of HIF-1. p38 MAPK can also cause the activation of SIAH2, which results in the ubiquitination and degradation of PHD3. Important in this model of HIF-1 activation are also phosphatases, in particular DUSP1 and DUSP2—enzymes that catalyze a reaction reverse to ERK MAPK and p38 MAPK. In chronic hypoxia, there is a decrease in DUSP2 expression but an increase in DUSP1, which is a mechanism for regulating HIF-1 activation.
Ijms 22 10701 g002
Figure 3. The hypoxia-induced mechanism of NF-κB activation. In hypoxia, NF-κB is an important factor in the increase in HIF-1 mRNA expression, which is activated when oxygen concentration is decreased. This process occurs through multiple pathways. Like HIF-1α, IKKβ activation is inhibited by hydroxylation by PHD1. In hypoxia, PHD1 activity is reduced, which enables IKKβ activation. NF-κB activation during hypoxia also involves ROS and calcium ion mobilization into the cytoplasm. These factors cause the ubiquitination of IKKγ/NEMO, which increases IKK activity. IκBα is SUMOylated, which decreases the activity of this inhibitor of the NF-κB activation pathway. Chronic hypoxia is also associated with the activation of kinases such as p38 MAPK, ERK MAPK, and Akt/PKB, which phosphorylate NF-κB and IKKβ, thus activating this transcription factor.
Figure 3. The hypoxia-induced mechanism of NF-κB activation. In hypoxia, NF-κB is an important factor in the increase in HIF-1 mRNA expression, which is activated when oxygen concentration is decreased. This process occurs through multiple pathways. Like HIF-1α, IKKβ activation is inhibited by hydroxylation by PHD1. In hypoxia, PHD1 activity is reduced, which enables IKKβ activation. NF-κB activation during hypoxia also involves ROS and calcium ion mobilization into the cytoplasm. These factors cause the ubiquitination of IKKγ/NEMO, which increases IKK activity. IκBα is SUMOylated, which decreases the activity of this inhibitor of the NF-κB activation pathway. Chronic hypoxia is also associated with the activation of kinases such as p38 MAPK, ERK MAPK, and Akt/PKB, which phosphorylate NF-κB and IKKβ, thus activating this transcription factor.
Ijms 22 10701 g003
Figure 4. The inhibition of the NF-κB pathway activation by HIF. Chronic hypoxia is associated with NF-κB activation, although there are also mechanisms that silence the proinflammatory response, such as an increase in PHD3 expression, which inhibits IKK activity. Additionally, there is an HIF-1 induced increase in the expression of IκBα, an inhibitor of NF-κB. The simultaneous activation of NF-κB and HIF causes these two transcription factors to compete for the coactivator p300.
Figure 4. The inhibition of the NF-κB pathway activation by HIF. Chronic hypoxia is associated with NF-κB activation, although there are also mechanisms that silence the proinflammatory response, such as an increase in PHD3 expression, which inhibits IKK activity. Additionally, there is an HIF-1 induced increase in the expression of IκBα, an inhibitor of NF-κB. The simultaneous activation of NF-κB and HIF causes these two transcription factors to compete for the coactivator p300.
Ijms 22 10701 g004
Figure 5. Effect of ROS in cycling hypoxia on the activation of HIF-1 and NF-κB. Cycling hypoxia induces the generation of ROS, which cause the activation of HIF-1 and NF-κB. In particular, ROS inactivate FIH and PHD, which results in increased stability of HIF-1α protein. ROS also activate PKA and mTOR, which phosphorylate HIF-1α and thus increase the stability of this protein and its accumulation in the cell. ROS also causes an increase in the expression of Trx1, which enhances the transcriptional activity of HIF-1.
Figure 5. Effect of ROS in cycling hypoxia on the activation of HIF-1 and NF-κB. Cycling hypoxia induces the generation of ROS, which cause the activation of HIF-1 and NF-κB. In particular, ROS inactivate FIH and PHD, which results in increased stability of HIF-1α protein. ROS also activate PKA and mTOR, which phosphorylate HIF-1α and thus increase the stability of this protein and its accumulation in the cell. ROS also causes an increase in the expression of Trx1, which enhances the transcriptional activity of HIF-1.
Ijms 22 10701 g005
Figure 6. Effect of cycling hypoxia on angiogenesis in cancer. Cycling hypoxia activates HIF-1 and NF-κB in the tumor cell. (a) This leads to increased production of VEGF-A, CCL2/MCP-1, CXCL1/GRO-α, CXCL8/IL-8, and PGE2. (b) Subsequently, CCL2/MCP-1, CXCL1/GRO-α and CXCL8/IL-8 induce recruitment of TAM and TAN to the tumor niche. Cells that possess pro-angiogenic properties. TAN secrete MMP-9 into the tumor microenvironment, whereas TAM secrete MMP-9 and VEGF-A. MMP-9 is a metalloproteinase that releases VEGF-A. PGE2 also increases the expression of proangiogenic factors. (c) VEGF-A, CCL2/MCP-1, CXCL1/GRO-α and CXCL8/IL-8 directly cause angiogenesis.
Figure 6. Effect of cycling hypoxia on angiogenesis in cancer. Cycling hypoxia activates HIF-1 and NF-κB in the tumor cell. (a) This leads to increased production of VEGF-A, CCL2/MCP-1, CXCL1/GRO-α, CXCL8/IL-8, and PGE2. (b) Subsequently, CCL2/MCP-1, CXCL1/GRO-α and CXCL8/IL-8 induce recruitment of TAM and TAN to the tumor niche. Cells that possess pro-angiogenic properties. TAN secrete MMP-9 into the tumor microenvironment, whereas TAM secrete MMP-9 and VEGF-A. MMP-9 is a metalloproteinase that releases VEGF-A. PGE2 also increases the expression of proangiogenic factors. (c) VEGF-A, CCL2/MCP-1, CXCL1/GRO-α and CXCL8/IL-8 directly cause angiogenesis.
Ijms 22 10701 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Korbecki, J.; Simińska, D.; Gąssowska-Dobrowolska, M.; Listos, J.; Gutowska, I.; Chlubek, D.; Baranowska-Bosiacka, I. Chronic and Cycling Hypoxia: Drivers of Cancer Chronic Inflammation through HIF-1 and NF-κB Activation: A Review of the Molecular Mechanisms. Int. J. Mol. Sci. 2021, 22, 10701. https://doi.org/10.3390/ijms221910701

AMA Style

Korbecki J, Simińska D, Gąssowska-Dobrowolska M, Listos J, Gutowska I, Chlubek D, Baranowska-Bosiacka I. Chronic and Cycling Hypoxia: Drivers of Cancer Chronic Inflammation through HIF-1 and NF-κB Activation: A Review of the Molecular Mechanisms. International Journal of Molecular Sciences. 2021; 22(19):10701. https://doi.org/10.3390/ijms221910701

Chicago/Turabian Style

Korbecki, Jan, Donata Simińska, Magdalena Gąssowska-Dobrowolska, Joanna Listos, Izabela Gutowska, Dariusz Chlubek, and Irena Baranowska-Bosiacka. 2021. "Chronic and Cycling Hypoxia: Drivers of Cancer Chronic Inflammation through HIF-1 and NF-κB Activation: A Review of the Molecular Mechanisms" International Journal of Molecular Sciences 22, no. 19: 10701. https://doi.org/10.3390/ijms221910701

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