Open Access

The role of hypoxia in inflammatory disease (Review)

  • Authors:
    • John Biddlestone
    • Daniel Bandarra
    • Sonia Rocha
  • View Affiliations

  • Published online on: January 27, 2015     https://doi.org/10.3892/ijmm.2015.2079
  • Pages: 859-869
  • Copyright: © Biddlestone et al. This is an open access article distributed under the terms of Creative Commons Attribution License [CC BY_NC 3.0].

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Mammals have developed evolutionarily conserved programs of transcriptional response to hypoxia and inflammation. These stimuli commonly occur together in vivo and there is significant crosstalk between the transcription factors that are classically understood to respond to either hypoxia or inflammation. This crosstalk can be used to modulate the overall response to environmental stress. Several common disease processes are characterised by aberrant transcriptional programs in response to environmental stress. In this review, we discuss the current understanding of the role of the hypoxia-responsive (hypoxia-inducible factor) and inflammatory (nuclear factor-κB) transcription factor families and their crosstalk in rheumatoid arthritis, inflammatory bowel disease and colorectal cancer, with relevance for future therapies for the management of these conditions.

1. Introduction

Oxygen (O2) constitutes 20.8% of the atmospheric air, and is the third-most abundant element in the universe, after hydrogen and helium. It is not only a key component of all major biomolecules of living organisms, but also a key constituent of inorganic compounds. Oxygen homeostasis is crucially important to maintain the survival of all vertebrate species (1). Therefore, organisms developed a way to coordinate the oxygen levels in the intracellular compartments in order to maintain homeostasis. When these mechanisms fail, and the intracellular concentration of oxygen decreases, a stress condition called hypoxia is created. Hypoxia can be defined as a condition lacking the necessary oxygen to meet metabolic requirements. The level at which this is reached will vary depending on the metabolic requirements of the cell. Hypoxia is a relevant physiological stress associated with many processes, such as adaptation to high altitudes or human diseases (e.g, cancer) (2). The hypoxia-inducible factors (HIFs) are a family of transcription factors whose levels are regulated in response to hypoxic stimuli, and when active can enact a transcriptional program that allows the cell to respond to the hypoxic environment.

Another important physiological stress is inflammation. Inflammation represents a protective attempt to eliminate pathogens and initiate the healing process of a wound. As in hypoxia, cells have evolved sophisticated mechanisms to control the inflammatory response to pathogens. A key element of these mechanisms is a family of transcription factors termed nuclear factor κ-light-chain-enhancer of activated B cells (NF-κB). NF-κB is composed of several family members that activate signalling pathways in response to a variety of stimuli (such as virus, bacteria or cytokines) which ultimately engage a complex transcriptional program, allowing the cell to respond to this environmental stress (3).

Several diseases, including rheumatoid arthritis (RA), inflammatory bowel disease and colorectal cancer (CRC) result from the deregulation of the hypoxia and inflammation pathways (46). Consequently, recent scientific research has been focussed on attempting to understand how these pathways are regulated, crosstalk and respond in disease. In this review, we describe the current understanding of the role of the HIF and NF-κB transcription factor families in response to hypoxia and inflammation and discuss their crosstalk in RA, inflammatory bowel disease and CRC, with relevance for future therapies for the management of these conditions.

2. The HIF system

The HIFs are a family of transcription factors that sense changes in environmental oxygen and orchestrate a transcriptional program, which forms an important part of the cellular response to the hypoxic environment. HIF-1 was first identified over 20 years ago through studies of erythropoietin gene expression (7). HIF is a heterodimeric transcription factor that consists of a constitutively expressed HIF-1β subunit and an O2-regulated HIF-α subunit (8). Three isoforms of HIF-α have been identified since these initial studies (HIF-1α, -2α and -3α) (Fig. 1A). The HIF-α isoforms are all characterized by the presence of basic helix-loop-helix (bHLH)-Per/ARNT/Sim (PAS) and oxygen-dependent degradation (ODD) domains (Fig. 1A). Both HIF-1α and HIF-2α have important cellular functions as transcription factors with some redundancy in their targets (9,10). HIF-2α protein shares sequence similarity and functional overlap with HIF-1α, but its distribution is restricted to certain cell types, and in some cases, it mediates distinct biological functions (11). HIF-3α is the most recently discovered isoform. The regulation of HIF-3α expression is complex in comparison to HIF-1α and HIF-2α with several splice variants that can function as a competitive inhibitors of the HIF-α-HIF-1β interaction (12,13), or by directly activating target genes in hypoxia that mediate both hypoxia dependent and independent functions. The role of HIF-3α in the cellular response to hypoxia remains an active area of study (14). Several splice variants of HIF-1β [also known as aryl hydrocarbon receptor nuclear translocator (ARNT)] have been identified (15,16). Though their exact functions are not known, at least one splice variant has been associated with poor prognosis in oestrogen receptor-negative breast cancer (17).

3. HIF regulation by οxygen

The regulation of the HIF-α subunits by oxygen occurs mainly at the post-transcriptional level, and is mediated by hydroxylation-dependent proteasomal degradation (Fig. 1B). In well-oxygenated cells, HIF-α is hydroxylated in its ODD. For HIF-1α this is at prolines (Pro)402 and Pro564 (18), whereas HIF-2α is hydroxylated at Pro405 and Pro531 (Fig. 1B) (19). Proline hydroxylation is catalysed by a class of dioxygenase enzymes called prolyl hydroxylases (PHDs). There are three known PHDs, 1–3, all of which have been shown to hydroxylate HIF-1α. PHD2 has a higher affinity for HIF1α, whereas PHD1 and PHD3 have higher affinity for HIF-2α (20,21). All PHDs require Fe2+ and α-ketoglutarate (α-KG) as co-factors for their catalytic activity and have an absolute requirement for molecular oxygen as a co-substrate, making their activity reduced in hypoxia (2225).

Prolyl-hydroxylation of HIF-α attracts the von Hippel-Lindau (vHL) tumour suppressor protein, which recruits the Elongin C-Elongin B-Cullin 2-E3-ubiquitin-ligase complex, leading to the Lys48-linked poly-ubiquitination and proteasomal degradation of HIF-1α (Fig. 1B) (2628). Interestingly, PHDs have also been shown to be able to sense amino acid availability through α-KG oscillations (29), and the centrosomal protein Cep192 has been described as a hydroxylation target for PHD1 (30), indicating an additional function for these enzymes as nutrient sensors and regulators of cell cycle progression. Both PKM2 and HCLK2 have also both recently been described as new hydroxylation targets for PHD3 (31,32).

In hypoxia the PHDs are inactive, or have reduced activity, since they require molecular oxygen as a cofactor. Under these conditions HIF-α is stabilized, can form a heterodimer with HIF-1β in the nucleus and bind to the consensus cis-acting hypoxia response element (HRE) nucleotide sequence 5′-RCGTG-3′, which is present within the enhancers and/or promoters of HIF target genes (Fig. 1B) (3335). HIF-α stabilisation therefore allows the cell to enact a transcriptional programme that is appropriate to the hypoxic environment (18) (Fig. 1B).

4. HIF target genes

The HIF heterodimer can regulate the expression of over 100 target genes involved in a broad range of physiological functions including: angiogenesis, erythropoiesis, metabolism, autophagy, apoptosis and other physiological responses to hypoxia (36). Canonical HIF signalling is based on the recognition of a putative HRE in the promoter or enhancer of the target gene that results in the recruitment of the HIF heterodimer and machinery required for transcription. Proteomics approaches have been used to identify protein changes in response to hypoxia in comparison with gene changes. Changes in just over 100 proteins in response to hypoxia have been identified (37,38). However, proteins identified represent both known and undescribed HIF targets, raising the possibility of HIF action outside of the conventional canonical pathway. Indeed, in addition to canonical signalling, there are various described mechanisms by which the stabilised HIF isoforms can influence the activity of other signalling pathways independent of the HIF heterodimer or a HRE. Non-canonical HIF signalling has been demonstrated to regulate aspects of Notch (39), c-Myc (40) and p53 (41) signalling.

5. Inflammation and the NF-κB pathway

Inflammation is a complex physiological process characterised by the activation of several coordinated signalling pathways in response to stress. Generally, the inflammatory response involves both anti- and pro-inflammatory mediators, given by the expression of small peptides (e.g., cytokines), glycoproteins (e.g., cluster of differentiation (CD)], and transcription factors, such as NF-κB.

NF-κB is considered the main pro-inflammatory family of transcription factors (4244). In mammalians, it is characterised as a family of five Rel-domain proteins; RelA, RelB, cRel, p100/p52 and p105/p50 (Fig. 2A). Interestingly, it has been shown that almost all combinations of homo- or hetero-dimers between the five NF-κB subunits are possible (45). This is important, not only because it gives an extra layer of complexity to the NF-κB system, but also because it gives specificity according to cellular context, stimuli or DNA sequences that are bound to the subunits (44,46). All the NF-κB subunits are characterised by a conserved 300-amino acid domain, the Rel homology domain (RHD), which is located in the N-terminus of the protein (Fig. 2A), and is responsible for dimerisation, and DNA binding. While RelA, RelB and cRel contain a C-terminal transactivation domain (TAD) (Fig. 2A), p105 and p100 contain Ankyrin-repeats motifs in their C-terminus (ANK) (Fig. 2A), responsible for the dimerisation with other subunits, and subsequent sequestration/inactivation in the cytoplasm (Fig. 2B).

There are distinct pathways for the activation of NF-κB, according to the stimulus, as well as the kinases and NF-κB subunits involved (3). The most common, and most well studied is the classical or canonical NF-κB pathway (Fig. 2B). In unstimulated cells, the NF-κB dimers remain inactive in the cytosol, bound to an inhibitory protein, inhibitor of NF-κB (IκB) (47). Upon stimulation, for example by the pro-inflammatory cytokine, tumour necrosis factor-α (TNF-α), the inhibitor of κB kinase (IκB kinase; IKK), is activated, and phosphorylates IκB. This leads to the degradation of IκB and the release/translocation of the NF-κB complex into the nucleus (48). In the nucleus, the activated NF-κB complex binds to specific 9–10 base pair DNA sequences (κB sites) to activate a complex regulatory network in response to a specific stimulus (49). The combination of different possible homo- and heterodimers, stimuli and cellular context leads to a myriad of possible outcomes, namely the activation or inhibition of apoptosis, cellular growth and carcinogenesis (50).

The NF-κB system is complex and is involved in multiple biological roles; it is thus expected that it is deregulated in many different diseases. NF-κB abnormal activation has been associated with several human diseases, such as inflammation-related diseases (inflammatory bowel disease and asthma), cancer (apoptosis suppression), viral infections (HIV) and genetic diseases (incontinentia pigmenti) (51).

6. Crosstalk between hypoxia and inflammation in disease

Hypoxia and inflammation are intimately linked. It has been reported that individuals with mountain sickness presented with increased inflammatory cytokines circulating in the blood (52). Additionally, healthy volunteers who have been exposed to a hypoxic environment for three nights in high altitudes (>3,400 meters), presented with high levels of the inflammatory cytokine, interleukin (IL)-6, in the blood (53). On the other hand, several inflammatory diseases, such as RA and inflammatory bowel disease, also exhibit areas of combined hypoxia and inflammation, which are usually associated with a poor prognosis of the disease (5457).

Hypoxia and inflammation are also connected at the molecular level (48,58,59). HIF (hypoxia) and NF-κB (inflammation) have been shown to have several common target genes, common regulators, and importantly, common stimuli (48). NF-κB activation has been shown to stabilise HIF-1α in hypoxia, and, together with HIF-1β, in inflammation (60,61). On the other hand, HIF-1α has been shown to repress NF-κB in vivo and in vitro under inflammatory conditions (59,62,63). The complexity of the combined response of HIF and NF-κB in hypoxia makes the crosstalk of these two pathways more intricate, and difficult to study. However, by developing a suitable inflammatory model, where the pathways can be controlled, as well as the conditions of the stimuli, these studies could provide very useful information that ultimately should be used to uncover new therapeutic strategies in a diverse range of diseases where hypoxia and inflammation are predominant features. In this review, the crosstalk between the main players induced in both inflammation and hypoxia in three clinical settings is addressed.

Hypoxia and inflammation crosstalk in RA

RA is a systemic autoimmune disorder characterised by chronic inflammation of the synovial membranes of joint tissues at multiple anatomical sites which ultimately leads to localised destruction and debilitating deformity (64,65). The RA joint synovium is characterised by both inflammatory and hypoxic regions (Fig. 3), which are highly infiltrated with lymphocytes (CD4+ T cells, and B cells), macrophages and macrophage-like and fibroblast-like synoviocytes (66). The molecular basis of RA is still poorly understood, mainly because RA is a heterogeneous disease composed of several possible treatment responses, and clinical manifestations (6769). These differences make RA difficult to treat, and further studies on the crosstalk between pathways involved in the disease are required.

The role of NF-κB in RA

The deregulation of several transcription factors, such as NF-κB, activator protein-1 (AP-1), and signal transducer and activator of transcription (STATs), has been strongly associated with the inflammatory setting of RA (7072). NF-κB, in particular, has been shown to be highly activated in the RA synovium (73,74). This is exceptionally important due to the major role of NF-κB in activating inflammatory responses, such as through the activation of the pro-inflammatory cytokine, TNF-α, or the chemokine, IL-8 (75). The activation of a coordinated and complex network of pro-inflammatory cytokines, chemokines, metalloproteases (MPPs) and metabolic proteins by NF-κB, leads to the activation of a positive feedback loop, enhancing the activation of more pro-inflammatory signals that ultimately results in chronic and persistent inflammation (Fig. 3) (75,76).

The role of HIF in RA

The HIF family of proteins are additional transcription factors with direct relevance to RA (77,78). Recently, HIF-1α was identified as a key player in RA, and therefore as a potential therapeutic target (79). HIF is important to coordinate the hypoxia response in the synovial tissue, and the deregulation or failure of that response leads to cellular dysfunction, and can ultimately lead to cell death (80). Furthermore, the intense hypoxic region in the synovial tissue (2–4%), activates a hypoxic response through HIF, which is involved in regulating several genes involved in apoptosis, vasomotor control, energy metabolism, and importantly, angiogenesis (Fig. 3) (16,48,8183).

Even though the role of HIF in RA has been firmly established, the contribution of each α-subunit remains poorly understood. Recently, HIF-2α was implicated as the essential catabolic regulator of inflammation in RA (78). In that study, the authors demonstrated that the overexpression of HIF-2α in joint tissues, but not HIF-1α, was sufficient to induce RA pathogenesis (78). The full contribution of the α-subunits to RA remains elusive. However, it seems clear that each α-subunit contributes differently to the progression of RA. HIF-1α plays a more anti-inflammatory role, whereas HIF-2α acts in a pro-inflammatory manner. What regulates this differential expression of the isoforms is still unknown. However, taking into consideration that NF-κB is the main activator of the HIF transcription factors, it would be interesting to understand whether NF-κB has any role in this HIF-1α to HIF-2α switch, and whether that would be dependent of the presence of hypoxia, inflammation, or both combined.

Inflammatory bowel disease (IBD)

The intestinal mucosa is exposed to steep hypoxic gradients (63) and is in a constant state of controlled inflammation, which is necessary to allow tolerance to otherwise harmless ingested dietary antigens (Fig. 4) (84). This fine balance is pathologically disturbed in inflammatory bowel disease (IBD); a relapsing-remitting progressive disorder of the gastrointestinal tract that comprises both Crohn’s and ulcerative colitis. The symptoms of IBD can range from mild to severe and include abdominal pain, intestinal bleeding, weight loss, fever and diarrhoea (85). The two IBD sub-types have different distribution patterns: ulcerative colitis is restricted to the colon, whereas Crohn’s colitis can affect any part of the GI tract. Both are thought to occur when inappropriate immunological activity in the intestinal mucosa results in epithelial barrier dysfunction leading to exposure of the mucosal immune system to luminal antigenic material and further cycles of inflammation and barrier dysfunction that underlie disease progression (86,87).

Role of the HIF system in IBD

Hypoxia has been found to play a role in IBD. Lower resting oxygen levels have been demonstrated in sections of IBD tissue compared to the controls using a 2-nitroimidazole based approach (63). In keeping with these observations, HIF-1α and HIF-2α activation has been associated with disease and increased vascular density in human specimens (88). The increased vascular density was subsequently demonstrated to be effected by vascular endothelial growth factor (VEGF), an established target of the HIF system (89). Compartmental analyses of the effects of hypoxia have been possible in murine models, where hypoxia has been shown to affect the epithelium, primarily during periods of inflammation (63). The colonic epithelium is the most hypoxic and HIF-active tissue layer because it is physically farthest away from the colonic vascular plexus and closest to the anoxic bowel lumen. This effect is exacerbated by oxygen consumption by luminal bacteria (90), and the presence of inflammatory mediators and lipopolysaccharide (LPS), which have been shown to regulate HIF activity (48).

In the context of IBD, HIF system activity is thought to be protective, acting through three mechanisms: i) inhibition of epithelial cell apoptosis; ii) enhanced expression of barrier-protective genes; and the iii) promotion of neutrophil apoptosis (Fig. 4) (86). Evidence of the anti-apoptotic effects of HIF has been demonstrated indirectly through experiments to investigate the role of the hydroxylase inhibitor, dimethyloxaloylglycine (DMOG), in colitis. Using a murine model of dextran sodium sulfate (DSS)-induced colitis, HIF stabilisation following treatment with DMOG has been shown to prevent apoptosis in a mechanism thought to be mediated by the anti-apoptotic protein, cIAP-2 (91). Recently, this effect has been specifically attributed to PHD1, since the homozygous loss of PHD1, but not PHD2 or PHD3, has been shown to be protective in the same murine model of DSS-induced colitis (92). This effect is most likely HIF-dependent, since the conditional knockout of HIF-1α in mouse intestinal epithelial cells has been shown to result in an enhanced susceptibility to the development of colitis (63).

In addition to its anti-apoptotic effects, the HIF system can protect against colitis through the expression of barrier-protective genes. Several HIF-dependent target genes have been proposed as mediators of this effect: CD55 (93), ecto-50 nucleotidase (94), A2B receptor (95) MUC-3 (96), intestinal trefoil factor (97), and P-glycoprotein (98) all play a role in the regulation of the intestinal mucosa barrier and have all been demonstrated to be regulated in a hypoxia-dependent manner.

There is also evidence of the differential effects of the HIF-α isoform in IBD. HIF-2α expression has been shown to be increased in colon tissues of mice after the induction of colitis. This was also observed in patients with ulcerative colitis or Crohn’s disease (62). Interestingly, in that study, while the loss of HIF-2α was associated with attenuated colonic inflammation, the overexpression of HIF-2α led to spontaneous colitis and increased inflammation.

Role of NF-κB in IBD

Other transcriptional programs are active in IBD in addition to those enacted by the HIF system. IBD is primarily an inflammatory pathology and NF-κB activity has been linked to its progression (5). A high degree of NF-κB induction has been demonstrated in intestinal macrophages and epithelial cells (99). In IBD, inflammatory cytokines can drive NF-κB activation, leading to the production of more inflammatory cytokines and potentiating further NF-κB activation (Fig. 4). NF-κB-induced TNF-α expression is one example of this type of positive feedback loop (100). Interestingly, NF-κB can have a dual role in IBD, potentiating inflammation in intestinal macrophages while protecting from inflammation in mucosal epithelial cells. Sharing interesting similarities to the effects of HIF activation, NF-κB signalling in intestinal epithelial cells has been shown to be protective against the development of colitis (101). Deletion of the NF-κB pathway in intestinal epithelial cells results in decreased expression of anti-apoptotic genes, such as Bcl-xL, and leads to reduced epithelial barrier function and increased susceptibility to colitis (102). Conditional knockout of NEMO and subsequent NF-κB inhibition has been shown to result in severe epithelial inflammation in a murine model (102). Similarly, epithelial cell-specific IKKβ deletion has been shown to result in the sustained production of pro-inflammatory Th1 cytokines and increased intestinal inflammation (103). Several treatments have been proposed to target NF-κB activity in IBD, including proteasome blockade, the administration of non-coding RNAs to interfere with NF-κB-DNA binding and anti-TNF-α immunotherapy. However, all have been met with significant systemic toxicity due to the broad role of NF-κB in multiple organs.

NF-κB-HIF crosstalk in IBD

Sharing similarities with the microenvironment of RA, in IBD both inflammation and hypoxia are present in the intestinal epithelium and contribute to disease progression (104). It is generally [but not universally (105)] understood that both NF-κB and HIF activity are protective in episodes of colitis (101). Significant crosstalk between these pathways has already been established, and it has been proposed that both pathways may act in concert to contribute to the epithelial barrier function of the colon in a process that is deregulated in IBD.

One example of this crosstalk is the regulation of apoptosis by both pathways (106,107). The caspase recruitment domain family, member 9 (CARD9) is understood to function as a molecular scaffold for the assembly of a BCL10 signalling complex that activates NF-κB (106), and has also been shown to be involved in the regulation of hypoxia-sensitive pathways (107). CARD9 therefore represents one point of crosstalk that may be important in the development of IBD as a promising target for further investigation.

Our laboratory and others have demonstrated NF-κB-dependent HIF-1α mRNA regulation (61,108). NF-κB can also regulate HIF signalling through IKKγ and HIF-2α, which increases HIF-2α transcriptional activity through interaction with cAMP response element-binding (CREB) binding protein (CBP)/p300 (109). Negative feedback through the NF-κB-dependent induction of the micro-RNA, miR-155, in response to LPS has been shown to target HIF-1α for silencing (110). Furthermore, our laboratory have recently demonstrated an evolutionarily conserved negative feedback mechanism through which HIF can regulate NF-κB in a mechanism that is dependent on the kinases, TAK-IKK and CDK6 (59).

CRC

CRC is a lethal disease affecting over 500,000 individuals annually (111). In contrast to the protective effects of HIF and NF-κB activity in IBD, both can play important roles in the development of colorectal malignancy. In CRC, the hypoxic milieu is similar to that of IBD but, critically, the cells are transformed to allow them to react differently to the activation of either system. In addition, chronic inflammation is a hallmark of cancer (112). The role of the HIF system and the role of NF-κB activity are considered below, and the significance of their crosstalk with respect to the development of CRC is examined.

Role of the HIF system in CRC

The role of the HIFs in cancer progression has long been appreciated due to their ability to promote angiogenesis through one of the principally identified HIF-1α target genes, VEGFA (113). However, it is becoming more evident that hypoxia and the HIF system can affect tumour growth through modulation of proliferation, apoptosis and epithelial to mesenchymal transition (EMT) (Fig. 5). Hypoxia and the subsequent HIF activation are generally understood to be prognostically bad and lead to tumour progression (114). In CRC, HIF-1α stabilisation has been shown to lead to a poor disease outcome. Shay et al (114) demonstrated that the inhibition of HIF signalling using acriflavine halted the progression of an autochthonous model of established colitis-associated colon cancer in immuno-competent mice. In their model, treatment with acriflavine was shown to decrease tumour number, size and advancement, in an effect thought to be mediated through the inhibition of HIF-dependent targets, such as VEGFA. These data provide a direct link between HIF-1α expression and tumour progression. However, HIF isoform activation can be antagonistic in the context of tumour progression. In contrast to the effect of high HIF-1α expression, high HIF-2α expression has recently been reported to prevent CRC progression (115). The antagonistic effects of HIF-1α and HIF-2α are important for the regulation of proliferation and apoptosis in cancer biology (40,82,116). The HIF system can affect proliferation through the regulation of cMyc. HIF-1α can promote cell cycle arrest by the direct opposition of c-Myc activity and the induction of p21 in CRC (116). Conversely, HIF-2α has been shown to promote proliferation in through its augmentation of cMyc function (40).

The HIF system can also affect apoptosis through the regulation of p53. p53 stability leads to apoptosis in somatic cells and it is frequently mutated in cancers in pursuit of immortality. HIF-1α has been shown to stabilise wild-type p53 via physical interaction through its ODD (41,117). As a form of negative feedback, p53 can promote the degradation of HIF-1α (118). The negative feedback of wild-type p53 on HIF-1α could explain the increased stability of HIF-1α in tumours that express mutant p53 which is incapable of degrading HIF-1α. The net result of the p53-HIF-1α interaction is increased apoptosis in damaged cells that are exposed to hypoxia (119). HIF-2α can inhibit p53 phosphorylation, resulting in a reduction in p53 pathway activity and the prevention of apoptosis in response to damaging stimuli (120). In addition to its role in the regulation of p53, HIF has been linked to the positive regulation of apoptosis through the control of several pro-apoptotic factors, including caspase-3, Fas and Fas ligand (121).

Hypoxia is a critical determinant of the motile and invasive phenotype of cancer cells. HIF activation is also important in the regulation of genes involved in EMT, including the direct regulation of the EMT-promoting transcription factors, Snail and Twist, which have both been described as direct targets of the HIF system (122124). EMT is a critical event in the induction of tumour metastasis (125). Notch has also been shown to mediate HIF-1α-dependent EMT (126).

Role of NF-κB in CRC

The role of NF-κB in CRC is an active area of study (101,127129). Inflammation is an important trigger in the establishment and development of CRC. Patients with long-standing IBD have an increased risk of developing CRC (127,128). In this context, NF-κB activation can promote tumourigenesis and CRC progression. In CRC, chronic inflammation results in sustained reactive oxygen species (ROS) production, leading to DNA damage (Fig. 5) (130). Treatment with non-steroidal anti-inflammatory drugs (NSAIDs) reduces the development of CRC in patients with IBD and hereditary CRC (131,132), and the inactivation of NF-κB signalling reduces the formation of inflammation-associated tumours (101,129). IL-6 has been shown to be important for the number and size of tumours formed in mice (133), and IKKβ conditional knockout mice have been shown to develop more numerous tumours (134).

As with the HIF system, the mechanism of NF-κB-induced tumourigenesis and progression can be multifactorial. The activation of the NF-κB pathway confers survival, proliferation, angiogenic and migratory advantages (Fig. 5) (112,135138); all of which are hallmarks of cancer (112). NF-κB activation can block apoptosis by regulating the anti-apoptosis proteins, such as inhibitor of apoptotic proteins (IAPs) (139), or by the inhibition of prolonged c-Jun N-terminal kinase (JNK) signalling, modulating the accumulation of ROS (140). Alternatively, NF-κB activation can enhance IL-2 production, which can activate Janus kinase 3 (Jak3) by autophosphorylation (141). Jak3 can activate STAT3. Jak3 and STAT3 over-activation has been observed in human colon cancer in vivo and in vitro, and shown to prevent apoptosis, leading to poor prognosis (142,143). In addition, NF-κB activation can affect proliferation and cell growth through the regulation of its target genes, cyclin D1 and cMyc (144146), and promote angiogenesis through the regulation of VEGF and IL-8 (136). Finally, NF-κB activation has been shown to affect the expression of matrix metalloprotease-9 (MMP-9), in murine colon adenocarcinoma cells (147), an important protein in the regulation of migration and invasion.

NF-κB-HIF crosstalk in CRC

The data presented above demonstrate a clear overlap between the effectors of the HIF and NF-κB systems in the establishment and development of CRC. Solid tumours are characterised by the presence of hypoxia, as well as inflammation (6). Potential points for crosstalk include the regulation of cMyc and p53 (Fig. 5). NF-κB interacts with the co-activators, p300 and CREB-binding protein, to inhibit p53 function. This effect is reinforced by the NF-κB-dependent upregulation of the p53 inhibitor, mouse double minute 2 (MDM2) (3,148) and is similar to that exerted on p53 by HIF-1α. The expression of NF-κB, HIF, VEGF and Bcl-3 has been shown to correlate with proliferation, angiogenesis, decreased survival and a poor clinical outcome (149,150). In addition, TNF-α has been shown to stabilise Snail and β-catenin in a process that requires the downregulation of glycogen synthase kinase-3β (GSK3β) by NF-κB and the activation of Akt cascades, resulting in the promotion of EMT (151). These data are clinically important since NF-κB and Twist have been associated with lymph node metastasis in patients with CRC (152). Interestingly, HIF has been shown to interact with both Snail and Twist, making this another potential point for crosstalk between the pathways (153).

In addition to the mechanisms outlined above, there is a complex interplay between HIF, NF-κB and adenomatous polyposis coli (APC), that appears to be important in CRC (Fig. 5). One of the earliest events in the development of CRC is loss of the APC gene. Our laboratory has recently reported a functional crosstalk between HIF-1α and APC at the transcriptional level (154). HIF activation represses APC expression, acting at its promoter to result in positive activation and proliferation through the Wnt/β-catenin signalling and the TCF-LEF pathway (155), reduction in genetic and microtubule stability and reductions in cell migration (6,156).

The repression of APC by HIF-1α is complicated by the fact that medium levels of β-catenin can induce NF-κB, resulting in positive feedback, and high levels of β-catenin inhibit NF-κB, resulting in negative feedback (6). Further studies are required to determine the functional significance of this interaction in vivo. However, it represents another exciting point of crosstalk with importance for CRC disease progression.

7. Conclusion

In this review, the current understanding of the mechanisms of the HIF and NF-κB systems has been discussed with specific reference to the crosstalk between these two stress-responsive pathways. This crosstalk is significant for many disease processes and its role in RA, inflammatory bowel disease and CRC has been discussed in detail. It is important to note that the crosstalk between these pathways has significance beyond pathological processes. For example, in healthy individuals who live at a high altitude, prolonged HIF activation can lead to reduced NF-κB activity, effectively dampening the immune response. Further studies in this area is required; however, it is interesting that the anecdotal evidence of increased H. pylori infection in Tibetan monks exists (157). Individuals with mountain sickness have presented with increased levels of inflammatory cytokines circulating in the blood (52). Another study demonstrated that healthy volunteers who spent three nights at high altitudes (>3,400 meters), presented with high levels of the inflammatory cytokine, IL-6 (53). This hypoxia-inflammation crosstalk is also relevant in the clinical context. It was shown that ischemia in organ grafts increased the risk of inflammation and, consequently, graft failure or organ rejection (158). Accurate systematic experimentation is important to determine the mechanisms of the crosstalk between these pathways since these findings may have an impact on multiple disease processes, apart from those discussed herein. These include diabetes and systemic sclerosis, where limb perfusion is not optimal, resulting in increased tissue breakdown in the absence of an appropriate inflammatory response, leading to an increased infection rate.

Acknowledgments

We would like to thank members of the S.R. laboratory for their helpful discussions. J.B. is a CR-UK clinical fellow, D.B. is funded by a Wellcome Trust ISSF award, the S.R. laboratory is funded by a CR-UK Senior Research Fellowship (C99667/A12918). This study was also supported by a Wellcome Trust Strategic Award (097945/B/11/Z).

Abbreviations:

IKK

inhibitor of κB kinase

NF-κB

nuclear factor-κB

HIF

hypoxia-inducible factor

ARNT

aryl hydrocarbon nuclear translocator

PHD

prolyl hydroxylase

vHL

von Hippel Lindau

TNF

tumour necrosis factor

References

1 

Semenza GL: Regulation of oxygen homeostasis by hypoxia-inducible factor 1. Physiology (Bethesda). 24:97–106. 2009. View Article : Google Scholar

2 

Semenza GL: HIF-1 and human disease: One highly involved factor. Genes Dev. 14:1983–1991. 2000.PubMed/NCBI

3 

Perkins ND: The diverse and complex roles of NF-κB subunits in cancer. Nat Rev Cancer. 12:121–132. 2012.PubMed/NCBI

4 

Thornton RD, Lane P, Borghaei RC, Pease EA, Caro J and Mochan E: Interleukin 1 induces hypoxia-inducible factor 1 in human gingival and synovial fibroblasts. Biochem J. 350:307–312. 2000. View Article : Google Scholar : PubMed/NCBI

5 

Taylor CT: Interdependent roles for hypoxia inducible factor and nuclear factor-kappaB in hypoxic inflammation. J Physiol. 586:4055–4059. 2008. View Article : Google Scholar : PubMed/NCBI

6 

Näthke I and Rocha S: Antagonistic crosstalk between APC and HIF-1α. Cell Cycle. 10:1545–1547. 2011. View Article : Google Scholar

7 

Semenza GL and Wang GL: A nuclear factor induced by hypoxia via de novo protein synthesis binds to the human erythropoietin gene enhancer at a site required for transcriptional activation. Mol Cell Biol. 12:5447–5454. 1992.PubMed/NCBI

8 

Wang GL, Jiang BH, Rue EA and Semenza GL: Hypoxia-inducible factor 1 is a basic-helix-loop-helix-PAS heterodimer regulated by cellular O2 tension. Proc Natl Acad Sci USA. 92:5510–5514. 1995. View Article : Google Scholar

9 

Carroll VA and Ashcroft M: Role of hypoxia-inducible factor (HIF)-1alpha versus HIF-2alpha in the regulation of HIF target genes in response to hypoxia, insulin-like growth factor-I, or loss of von Hippel-Lindau function: Implications for targeting the HIF pathway. Cancer Res. 66:6264–6270. 2006. View Article : Google Scholar : PubMed/NCBI

10 

Zhou J, Schmid T, Schnitzer S and Brüne B: Tumor hypoxia and cancer progression. Cancer Lett. 237:10–21. 2006. View Article : Google Scholar

11 

Patel SA and Simon MC: Biology of hypoxia-inducible factor-2alpha in development and disease. Cell Death Differ. 15:628–634. 2008. View Article : Google Scholar : PubMed/NCBI

12 

Makino Y, Cao R, Svensson K, Bertilsson G, Asman M, Tanaka H, Cao Y, Berkenstam A and Poellinger L: Inhibitory PAS domain protein is a negative regulator of hypoxia-inducible gene expression. Nature. 414:550–554. 2001. View Article : Google Scholar : PubMed/NCBI

13 

Yamashita T, Ohneda O, Nagano M, et al: Abnormal heart development and lung remodeling in mice lacking the hypoxia-inducible factor-related basic helix-loop-helix PAS protein NEPAS. Mol Cell Biol. 28:1285–1297. 2008. View Article : Google Scholar :

14 

Zhang P, Yao Q, Lu L, Li Y, Chen PJ and Duan C: Hypoxia-inducible factor 3 is an oxygen-dependent transcription activator and regulates a distinct transcriptional response to hypoxia. Cell Rep. 6:1110–1121. 2014. View Article : Google Scholar : PubMed/NCBI

15 

Bárdos JI and Ashcroft M: Negative and positive regulation of HIF-1: A complex network. Biochim Biophys Acta. 1755:107–120. 2005.PubMed/NCBI

16 

Rocha S: Gene regulation under low oxygen: Holding your breath for transcription. Trends Biochem Sci. 32:389–397. 2007. View Article : Google Scholar : PubMed/NCBI

17 

Qin C, Wilson C, Blancher C, Taylor M, Safe S and Harris AL: Association of ARNT splice variants with estrogen receptor-negative breast cancer, poor induction of vascular endothelial growth factor under hypoxia, and poor prognosis. Clin Cancer Res. 7:818–823. 2001.PubMed/NCBI

18 

Kaelin WG Jr and Ratcliffe PJ: Oxygen sensing by metazoans: The central role of the HIF hydroxylase pathway. Mol Cell. 30:393–402. 2008. View Article : Google Scholar : PubMed/NCBI

19 

Haase VH: Renal cancer: Oxygen meets metabolism. Exp Cell Res. 318:1057–1067. 2012. View Article : Google Scholar : PubMed/NCBI

20 

Berra E, Benizri E, Ginouvès A, Volmat V, Roux D and Pouysségur J: HIF prolyl-hydroxylase 2 is the key oxygen sensor setting low steady-state levels of HIF-1alpha in normoxia. EMBO J. 22:4082–4090. 2003. View Article : Google Scholar : PubMed/NCBI

21 

Appelhoff RJ, Tian YM, Raval RR, Turley H, Harris AL, Pugh CW, Ratcliffe PJ and Gleadle JM: Differential function of the prolyl hydroxylases PHD1, PHD2, and PHD3 in the regulation of hypoxia-inducible factor. J Biol Chem. 279:38458–38465. 2004. View Article : Google Scholar : PubMed/NCBI

22 

Epstein AC, Gleadle JM, McNeill LA, et al: C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell. 107:43–54. 2001. View Article : Google Scholar : PubMed/NCBI

23 

Fandrey J, Gorr TA and Gassmann M: Regulating cellular oxygen sensing by hydroxylation. Cardiovasc Res. 71:642–651. 2006. View Article : Google Scholar : PubMed/NCBI

24 

Bruegge K, Jelkmann W and Metzen E: Hydroxylation of hypoxia-inducible transcription factors and chemical compounds targeting the HIF-alpha hydroxylases. Curr Med Chem. 14:1853–1862. 2007. View Article : Google Scholar : PubMed/NCBI

25 

Frede S, Stockmann C, Freitag P and Fandrey J: Bacterial lipopolysaccharide induces HIF-1 activation in human monocytes via p44/42 MAPK and NF-kappaB. Biochem J. 396:517–527. 2006. View Article : Google Scholar : PubMed/NCBI

26 

Ivan M, Kondo K, Yang H, Kim W, Valiando J, Ohh M, Salic A, Asara JM, Lane WS and Kaelin WG Jr: HIFalpha targeted for VHL-mediated destruction by proline hydroxylation: Implications for O2 sensing. Science. 292:464–468. 2001. View Article : Google Scholar : PubMed/NCBI

27 

Jaakkola P, Mole DR, Tian YM, et al: Targeting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science. 292:468–472. 2001. View Article : Google Scholar : PubMed/NCBI

28 

Yu F, White SB, Zhao Q and Lee FS: HIF-1alpha binding to VHL is regulated by stimulus-sensitive proline hydroxylation. Proc Natl Acad Sci USA. 98:9630–9635. 2001. View Article : Google Scholar : PubMed/NCBI

29 

Durán RV, MacKenzie ED, Boulahbel H, Frezza C, Heiserich L, Tardito S, Bussolati O, Rocha S, Hall MN and Gottlieb E: HIF-independent role of prolyl hydroxylases in the cellular response to amino acids. Oncogene. 32:4549–4556. 2013. View Article : Google Scholar :

30 

Moser SC, Bensaddek D, Ortmann B, Maure JF, Mudie S, Blow JJ, Lamond AI, Swedlow JR and Rocha S: PHD1 links cell-cycle progression to oxygen sensing through hydroxylation of the centrosomal protein Cep192. Dev Cell. 26:381–392. 2013. View Article : Google Scholar : PubMed/NCBI

31 

Luo W, Hu H, Chang R, Zhong J, Knabel M, O’Meally R, Cole RN, Pandey A and Semenza GL: Pyruvate kinase M2 is a PHD3-stimulated coactivator for hypoxia-inducible factor 1. Cell. 145:732–744. 2011. View Article : Google Scholar : PubMed/NCBI

32 

Xie L, Pi X, Mishra A, Fong G, Peng J and Patterson C: PHD3-dependent hydroxylation of HCLK2 promotes the DNA damage response. J Clin Invest. 122:2827–2836. 2012. View Article : Google Scholar : PubMed/NCBI

33 

Pugh CW, Tan CC, Jones RW and Ratcliffe PJ: Functional analysis of an oxygen-regulated transcriptional enhancer lying 3′ to the mouse erythropoietin gene. Proc Natl Acad Sci USA. 88:10553–10557. 1991. View Article : Google Scholar

34 

Semenza GL, Jiang BH, Leung SW, Passantino R, Concordet JP, Maire P and Giallongo A: Hypoxia response elements in the aldolase A, enolase 1, and lactate dehydrogenase A gene promoters contain essential binding sites for hypoxia-inducible factor 1. J Biol Chem. 271:32529–32537. 1996. View Article : Google Scholar : PubMed/NCBI

35 

Schödel J, Oikonomopoulos S, Ragoussis J, Pugh CW, Ratcliffe PJ and Mole DR: High-resolution genome-wide mapping of HIF-binding sites by ChIP-seq. Blood. 117:e207–e217. 2011. View Article : Google Scholar : PubMed/NCBI

36 

Semenza GL: Regulation of cancer cell metabolism by hypoxia-inducible factor 1. Semin Cancer Biol. 19:12–16. 2009. View Article : Google Scholar

37 

Han YH, Xia L, Song LP, Zheng Y, Chen WL, Zhang L, Huang Y, Chen GQ and Wang LS: Comparative proteomic analysis of hypoxia-treated and untreated human leukemic U937 cells. Proteomics. 6:3262–3274. 2006. View Article : Google Scholar : PubMed/NCBI

38 

Djidja MC, Chang J, Hadjiprocopis A, et al: Identification of hypoxia-regulated proteins using MALDI-mass spectrometry imaging combined with quantitative proteomics. J Proteome Res. 13:2297–2313. 2014. View Article : Google Scholar : PubMed/NCBI

39 

Gustafsson MV, Zheng X, Pereira T, Gradin K, Jin S, Lundkvist J, Ruas JL, Poellinger L, Lendahl U and Bondesson M: Hypoxia requires notch signaling to maintain the undifferentiated cell state. Dev Cell. 9:617–628. 2005. View Article : Google Scholar : PubMed/NCBI

40 

Gordan JD, Bertout JA, Hu CJ, Diehl JA and Simon MC: HIF-2alpha promotes hypoxic cell proliferation by enhancing c-myc transcriptional activity. Cancer Cell. 11:335–347. 2007. View Article : Google Scholar : PubMed/NCBI

41 

An WG, Kanekal M, Simon MC, Maltepe E, Blagosklonny MV and Neckers LM: Stabilization of wild-type p53 by hypoxia-inducible factor 1alpha. Nature. 392:405–408. 1998. View Article : Google Scholar : PubMed/NCBI

42 

Perkins ND: Integrating cell-signalling pathways with NF-kappaB and IKK function. Nat Rev Mol Cell Biol. 8:49–62. 2007. View Article : Google Scholar

43 

No authors listed. Celebrating 25 years of NF-κB. Nat Immunol. 12:6812011. View Article : Google Scholar

44 

Campbell KJ and Perkins ND: Regulation of NF-kappaB function. Biochem Soc Symp. 73:165–180. 2006.PubMed/NCBI

45 

Wong D, Teixeira A, Oikonomopoulos S, et al: Extensive characterization of NF-κB binding uncovers non-canonical motifs and advances the interpretation of genetic functional traits. Genome Biol. 12:R702011. View Article : Google Scholar

46 

Gilmore TD: The Rel/NF-kappaB signal transduction pathway: Introduction. Oncogene. 18:6842–6844. 1999. View Article : Google Scholar : PubMed/NCBI

47 

Chen F, Castranova V, Shi X and Demers LM: New insights into the role of nuclear factor-kappaB, a ubiquitous transcription factor in the initiation of diseases. Clin Chem. 45:7–17. 1999.PubMed/NCBI

48 

Bandarra DR and Rocha S: A tale of two transcription factors: NF-κB and HIF crosstalk. OA Mol Cell Biol. 1:62013. View Article : Google Scholar

49 

Gilmore TD: Introduction to NF-kappaB: Players, pathways, perspectives. Oncogene. 25:6680–6684. 2006. View Article : Google Scholar : PubMed/NCBI

50 

Perkins ND and Gilmore TD: Good cop, bad cop: The different faces of NF-kappaB. Cell Death Differ. 13:759–772. 2006. View Article : Google Scholar : PubMed/NCBI

51 

Aggarwal BB, Takada Y, Shishodia S, Gutierrez AM, Oommen OV, Ichikawa H, Baba Y and Kumar A: Nuclear transcription factor NF-kappa B: Role in biology and medicine. Indian J Exp Biol. 42:341–353. 2004.PubMed/NCBI

52 

Hackett PH and Roach RC: High-altitude illness. N Engl J Med. 345:107–114. 2001. View Article : Google Scholar : PubMed/NCBI

53 

Hartmann G, Tschöp M, Fischer R, Bidlingmaier C, Riepl R, Tschöp K, Hautmann H, Endres S and Toepfer M: High altitude increases circulating interleukin-6, interleukin-1 receptor antagonist and C-reactive protein. Cytokine. 12:246–252. 2000. View Article : Google Scholar : PubMed/NCBI

54 

Kim HL, Cho YS, Choi H, Chun YS, Lee ZH and Park JW: Hypoxia-inducible factor 1alpha is deregulated by the serum of rats with adjuvant-induced arthritis. Biochem Biophys Res Commun. 378:123–128. 2009. View Article : Google Scholar

55 

Boyd HK, Lappin TR and Bell AL: Evidence for impaired erythropoietin response to anaemia in rheumatoid disease. Br J Rheumatol. 30:255–259. 1991. View Article : Google Scholar : PubMed/NCBI

56 

Grenz A, Clambey E and Eltzschig HK: Hypoxia signaling during intestinal ischemia and inflammation. Curr Opin Crit Care. 18:178–185. 2012. View Article : Google Scholar : PubMed/NCBI

57 

Eltzschig HK, Sitkovsky MV and Robson SC: Purinergic signaling during inflammation. N Engl J Med. 367:2322–2333. 2012. View Article : Google Scholar : PubMed/NCBI

58 

Bandarra D, Biddlestone J, Mudie S, Muller HA and Rocha S: Hypoxia activates IKK-NF-κB and the immune response in Drosophila melanogaster. Biosci Rep. 34:342014. View Article : Google Scholar

59 

Bandarra D, Biddlestone J, Mudie S, Muller HA and Rocha S: HIF-1α restricts NF-κB dependent gene expression to control innate immunity signals. Dis Model Mech. Dec 15–2014.Epub ahead of print.

60 

van Uden P, Kenneth NS, Webster R, Müller HA, Mudie S and Rocha S: Evolutionary conserved regulation of HIF-1β by NF-κB. PLoS Genet. 7:e10012852011. View Article : Google Scholar

61 

van Uden P, Kenneth NS and Rocha S: Regulation of hypoxia-inducible factor-1alpha by NF-kappaB. Biochem J. 412:477–484. 2008. View Article : Google Scholar : PubMed/NCBI

62 

Xue X, Ramakrishnan S, Anderson E, Taylor M, Zimmermann EM, Spence JR, Huang S, Greenson JK and Shah YM: Endothelial PAS domain protein 1 activates the inflammatory response in the intestinal epithelium to promote colitis in mice. Gastroenterology. 145:831–841. 2013. View Article : Google Scholar : PubMed/NCBI

63 

Karhausen J, Furuta GT, Tomaszewski JE, Johnson RS, Colgan SP and Haase VH: Epithelial hypoxia-inducible factor-1 is protective in murine experimental colitis. J Clin Invest. 114:1098–1106. 2004. View Article : Google Scholar : PubMed/NCBI

64 

Sewell KL and Trentham DE: Pathogenesis of rheumatoid arthritis. Lancet. 341:283–286. 1993. View Article : Google Scholar : PubMed/NCBI

65 

Al-Shukaili AK and Al-Jabri AA: Rheumatoid arthritis, cytokines and hypoxia. What is the link. Saudi Med J. 27:1642–1649. 2006.PubMed/NCBI

66 

Gaber T, Dziurla R, Tripmacher R, Burmester GR and Buttgereit F: Hypoxia inducible factor (HIF) in rheumatology: Low O2! See what HIF can do. Ann Rheum Dis. 64:971–980. 2005. View Article : Google Scholar : PubMed/NCBI

67 

Hueber W, Kidd BA, Tomooka BH, et al: Antigen microarray profiling of autoantibodies in rheumatoid arthritis. Arthritis Rheum. 52:2645–2655. 2005. View Article : Google Scholar : PubMed/NCBI

68 

van Baarsen LG, Wijbrandts CA, Rustenburg F, Cantaert T, van der Pouw Kraan TC, Baeten DL, Dijkmans BA, Tak PP and Verweij CL: Regulation of IFN response gene activity during infliximab treatment in rheumatoid arthritis is associated with clinical response to treatment. Arthritis Res Ther. 12:R112010. View Article : Google Scholar : PubMed/NCBI

69 

van Wietmarschen HA, Dai W, van der Kooij AJ, et al: Characterization of rheumatoid arthritis subtypes using symptom profiles, clinical chemistry and metabolomics measurements. PLoS One. 7:e443312012. View Article : Google Scholar : PubMed/NCBI

70 

Sweeney SE and Firestein GS: Signal transduction in rheumatoid arthritis. Curr Opin Rheumatol. 16:231–237. 2004. View Article : Google Scholar : PubMed/NCBI

71 

Morel J and Berenbaum F: Signal transduction pathways: new targets for treating rheumatoid arthritis. Joint Bone Spine. 71:503–510. 2004. View Article : Google Scholar : PubMed/NCBI

72 

Firestein GS and Manning AM: Signal transduction and transcription factors in rheumatic disease. Arthritis Rheum. 42:609–621. 1999. View Article : Google Scholar : PubMed/NCBI

73 

Benito MJ, Murphy E, Murphy EP, van den Berg WB, FitzGerald O and Bresnihan B: Increased synovial tissue NF-kappa B1 expression at sites adjacent to the cartilage-pannus junction in rheumatoid arthritis. Arthritis Rheum. 50:1781–1787. 2004. View Article : Google Scholar : PubMed/NCBI

74 

Handel ML, McMorrow LB and Gravallese EM: Nuclear factor-kappa B in rheumatoid synovium. Localization of p50 and p65. Arthritis Rheum. 38:1762–1770. 1995. View Article : Google Scholar : PubMed/NCBI

75 

Müller-Ladner U, Pap T, Gay RE, Neidhart M and Gay S: Mechanisms of disease: The molecular and cellular basis of joint destruction in rheumatoid arthritis. Nat Clin Pract Rheumatol. 1:102–110. 2005. View Article : Google Scholar

76 

Simmonds RE and Foxwell BM: Signalling, inflammation and arthritis: NF-kappaB and its relevance to arthritis and inflammation. Rheumatology (Oxford). 47:584–590. 2008. View Article : Google Scholar

77 

Westra J, Molema G and Kallenberg CG: Hypoxia-inducible factor-1 as regulator of angiogenesis in rheumatoid arthritis -therapeutic implications. Curr Med Chem. 17:254–263. 2010. View Article : Google Scholar

78 

Ryu JH, Chae CS, Kwak JS, et al: Hypoxia-inducible factor-2α is an essential catabolic regulator of inflammatory rheumatoid arthritis. PLoS Biol. 12:e10018812014. View Article : Google Scholar

79 

Hu F, Shi L, Mu R, et al: Hypoxia-inducible factor-1α and interleukin 33 form a regulatory circuit to perpetuate the inflammation in rheumatoid arthritis. PLoS One. 8:e726502013. View Article : Google Scholar

80 

Brouwer E, Gouw AS, Posthumus MD, van Leeuwen MA, Boerboom AL, Bijzet J, Bos R, Limburg PC, Kallenberg CG and Westra J: Hypoxia inducible factor-1-alpha (HIF-1alpha) is related to both angiogenesis and inflammation in rheumatoid arthritis. Clin Exp Rheumatol. 27:945–951. 2009.

81 

Muz B, Khan MN, Kiriakidis S and Paleolog EM: Hypoxia. The role of hypoxia and HIF-dependent signalling events in rheumatoid arthritis. Arthritis Res Ther. 11:2012009. View Article : Google Scholar : PubMed/NCBI

82 

Moniz S, Biddlestone J and Rocha S: Grow2: The HIF system, energy homeostasis and the cell cycle. Histol Histopathol. 29:589–600. 2014.PubMed/NCBI

83 

Kenneth NS and Rocha S: Regulation of gene expression by hypoxia. Biochem J. 414:19–29. 2008. View Article : Google Scholar : PubMed/NCBI

84 

Poonam P: The biology of oral tolerance and issues related to oral vaccine design. Curr Pharm Des. 13:2001–2007. 2007. View Article : Google Scholar : PubMed/NCBI

85 

Podolsky DK: Inflammatory bowel disease. N Engl J Med. 347:417–429. 2002. View Article : Google Scholar : PubMed/NCBI

86 

Cummins EP, Doherty GA and Taylor CT: Hydroxylases as therapeutic targets in inflammatory bowel disease. Lab Invest. 93:378–383. 2013. View Article : Google Scholar : PubMed/NCBI

87 

Abraham C and Cho JH: Inflammatory bowel disease. N Engl J Med. 361:2066–2078. 2009. View Article : Google Scholar : PubMed/NCBI

88 

Giatromanolaki A, Sivridis E, Maltezos E, Papazoglou D, Simopoulos C, Gatter KC, Harris AL and Koukourakis MI: Hypoxia inducible factor 1alpha and 2alpha overexpression in inflammatory bowel disease. J Clin Pathol. 56:209–213. 2003. View Article : Google Scholar : PubMed/NCBI

89 

Danese S, Dejana E and Fiocchi C: Immune regulation by microvascular endothelial cells: Directing innate and adaptive immunity, coagulation, and inflammation. J Immunol. 178:6017–6022. 2007. View Article : Google Scholar : PubMed/NCBI

90 

Werth N, Beerlage C, Rosenberger C, et al: Activation of hypoxia inducible factor 1 is a general phenomenon in infections with human pathogens. PLoS One. 5:e115762010. View Article : Google Scholar : PubMed/NCBI

91 

Cummins EP, Seeballuck F, Keely SJ, Mangan NE, Callanan JJ, Fallon PG and Taylor CT: The hydroxylase inhibitor dimethyloxalylglycine is protective in a murine model of colitis. Gastroenterology. 134:156–165. 2008. View Article : Google Scholar : PubMed/NCBI

92 

Tambuwala MM, Cummins EP, Lenihan CR, et al: Loss of prolyl hydroxylase-1 protects against colitis through reduced epithelial cell apoptosis and increased barrier function. Gastroenterology. 139:2093–2101. 2010. View Article : Google Scholar : PubMed/NCBI

93 

Louis NA, Hamilton KE, Kong T and Colgan SP: HIF-dependent induction of apical CD55 coordinates epithelial clearance of neutrophils. FASEB J. 19:950–959. 2005. View Article : Google Scholar : PubMed/NCBI

94 

Synnestvedt K, Furuta GT, Comerford KM, Louis N, Karhausen J, Eltzschig HK, Hansen KR, Thompson LF and Colgan SP: Ecto-5′-nucleotidase (CD73) regulation by hypoxia-inducible factor-1 mediates permeability changes in intestinal epithelia. J Clin Invest. 110:993–1002. 2002. View Article : Google Scholar : PubMed/NCBI

95 

Kong T, Westerman KA, Faigle M, Eltzschig HK and Colgan SP: HIF-dependent induction of adenosine A2B receptor in hypoxia. FASEB J. 20:2242–2250. 2006. View Article : Google Scholar : PubMed/NCBI

96 

Louis NA, Hamilton KE, Canny G, Shekels LL, Ho SB and Colgan SP: Selective induction of mucin-3 by hypoxia in intestinal epithelia. J Cell Biochem. 99:1616–1627. 2006. View Article : Google Scholar : PubMed/NCBI

97 

Furuta GT, Turner JR, Taylor CT, Hershberg RM, Comerford K, Narravula S, Podolsky DK and Colgan SP: Hypoxia-inducible factor 1-dependent induction of intestinal trefoil factor protects barrier function during hypoxia. J Exp Med. 193:1027–1034. 2001. View Article : Google Scholar : PubMed/NCBI

98 

Comerford KM, Wallace TJ, Karhausen J, Louis NA, Montalto MC and Colgan SP: Hypoxia-inducible factor-1-dependent regulation of the multidrug resistance (MDR1) gene. Cancer Res. 62:3387–3394. 2002.PubMed/NCBI

99 

Neurath MF, Pettersson S, Meyer zum Büschenfelde KH and Strober W: Local administration of antisense phosphorothioate oligonucleotides to the p65 subunit of NF-kappa B abrogates established experimental colitis in mice. Nat Med. 2:998–1004. 1996. View Article : Google Scholar : PubMed/NCBI

100 

Holtmann MH and Neurath MF: Differential TNF-signaling in chronic inflammatory disorders. Curr Mol Med. 4:439–444. 2004. View Article : Google Scholar : PubMed/NCBI

101 

Greten FR, Eckmann L, Greten TF, Park JM, Li ZW, Egan LJ, Kagnoff MF and Karin M: IKKbeta links inflammation and tumorigenesis in a mouse model of colitis-associated cancer. Cell. 118:285–296. 2004. View Article : Google Scholar : PubMed/NCBI

102 

Pasparakis M: IKK/NF-kappaB signaling in intestinal epithelial cells controls immune homeostasis in the gut. Mucosal Immunol. 1(Suppl 1): S54–S57. 2008. View Article : Google Scholar : PubMed/NCBI

103 

Zaph C, Troy AE, Taylor BC, et al: Epithelial-cell-intrinsic IKK-beta expression regulates intestinal immune homeostasis. Nature. 446:552–556. 2007. View Article : Google Scholar : PubMed/NCBI

104 

Hauser CJ, Locke RR, Kao HW, Patterson J and Zipser RD: Visceral surface oxygen tension in experimental colitis in the rabbit. J Lab Clin Med. 112:68–71. 1988.PubMed/NCBI

105 

Shah YM, Ito S, Morimura K, et al: Hypoxia-inducible factor augments experimental colitis through an MIF-dependent inflammatory signaling cascade. Gastroenterology. 134:2036–2048. 2048 e2031–2033. 2008. View Article : Google Scholar : PubMed/NCBI

106 

Hara H and Saito T: CARD9 versus CARMA1 in innate and adaptive immunity. Trends Immunol. 30:234–242. 2009. View Article : Google Scholar : PubMed/NCBI

107 

Yang H, Minamishima YA, Yan Q, Schlisio S, Ebert BL, Zhang X, Zhang L, Kim WY, Olumi AF and Kaelin WG Jr: pVHL acts as an adaptor to promote the inhibitory phosphorylation of the NF-kappaB agonist Card9 by CK2. Mol Cell. 28:15–27. 2007. View Article : Google Scholar : PubMed/NCBI

108 

Rius J, Guma M, Schachtrup C, Akassoglou K, Zinkernagel AS, Nizet V, Johnson RS, Haddad GG and Karin M: NF-kappaB links innate immunity to the hypoxic response through transcriptional regulation of HIF-1alpha. Nature. 453:807–811. 2008. View Article : Google Scholar : PubMed/NCBI

109 

Bracken CP, Whitelaw ML and Peet DJ: Activity of hypoxia-inducible factor 2alpha is regulated by association with the NF-kappaB essential modulator. J Biol Chem. 280:14240–14251. 2005. View Article : Google Scholar : PubMed/NCBI

110 

O’Connell RM, Rao DS, Chaudhuri AA, Boldin MP, Taganov KD, Nicoll J, Paquette RL and Baltimore D: Sustained expression of microRNA-155 in hematopoietic stem cells causes a myeloproliferative disorder. J Exp Med. 205:585–594. 2008. View Article : Google Scholar

111 

Jemal A, Bray F, Center MM, Ferlay J, Ward E and Forman D: Global cancer statistics. CA Cancer J Clin. 61:69–90. 2011. View Article : Google Scholar : PubMed/NCBI

112 

Hanahan D and Weinberg RA: Hallmarks of cancer: The next generation. Cell. 144:646–674. 2011. View Article : Google Scholar : PubMed/NCBI

113 

Tsuzuki Y, Fukumura D, Oosthuyse B, Koike C, Carmeliet P and Jain RK: Vascular endothelial growth factor (VEGF) modulation by targeting hypoxia-inducible factor-1alpha--> hypoxia response element--> VEGF cascade differentially regulates vascular response and growth rate in tumors. Cancer Res. 60:6248–6252. 2000.PubMed/NCBI

114 

Shay JE, Imtiyaz HZ, Sivanand S, et al: Inhibition of hypoxia-inducible factors limits tumor progression in a mouse model of colorectal cancer. Carcinogenesis. 35:1067–1077. 2014. View Article : Google Scholar : PubMed/NCBI

115 

Rawluszko-Wieczorek AA, Horbacka K, Krokowicz P, Misztal M and Jagodzinski PP: Prognostic potential of DNA methylation and transcript levels of HIF1A and EPAS1 in colorectal cancer. Mol Cancer Res. 12:1112–1127. 2014. View Article : Google Scholar : PubMed/NCBI

116 

Koshiji M, Kageyama Y, Pete EA, Horikawa I, Barrett JC and Huang LE: HIF-1alpha induces cell cycle arrest by functionally counteracting Myc. EMBO J. 23:1949–1956. 2004. View Article : Google Scholar : PubMed/NCBI

117 

Sánchez-Puig N, Veprintsev DB and Fersht AR: Binding of natively unfolded HIF-1alpha ODD domain to p53. Mol Cell. 17:11–21. 2005. View Article : Google Scholar : PubMed/NCBI

118 

Ravi R, Mookerjee B, Bhujwalla ZM, Sutter CH, Artemov D, Zeng Q, Dillehay LE, Madan A, Semenza GL and Bedi A: Regulation of tumor angiogenesis by p53-induced degradation of hypoxia-inducible factor 1alpha. Genes Dev. 14:34–44. 2000.PubMed/NCBI

119 

Moeller BJ, Dreher MR, Rabbani ZN, Schroeder T, Cao Y, Li CY and Dewhirst MW: Pleiotropic effects of HIF-1 blockade on tumor radiosensitivity. Cancer Cell. 8:99–110. 2005. View Article : Google Scholar : PubMed/NCBI

120 

Bertout JA, Majmundar AJ, Gordan JD, Lam JC, Ditsworth D, Keith B, Brown EJ, Nathanson KL and Simon MC: HIF2alpha inhibition promotes p53 pathway activity, tumor cell death, and radiation responses. Proc Natl Acad Sci USA. 106:14391–14396. 2009. View Article : Google Scholar : PubMed/NCBI

121 

Volm M and Koomägi R: Hypoxia-inducible factor (HIF-1) and its relationship to apoptosis and proliferation in lung cancer. Anticancer Res. 20:1527–1533. 2000.PubMed/NCBI

122 

Evans AJ, Russell RC, Roche O, et al: VHL promotes E2 box-dependent E-cadherin transcription by HIF-mediated regulation of SIP1 and snail. Mol Cell Biol. 27:157–169. 2007. View Article : Google Scholar :

123 

Yang MH, Wu MZ, Chiou SH, Chen PM, Chang SY, Liu CJ, Teng SC and Wu KJ: Direct regulation of TWIST by HIF-1alpha promotes metastasis. Nat Cell Biol. 10:295–305. 2008. View Article : Google Scholar : PubMed/NCBI

124 

Gort EH, van Haaften G, Verlaan I, et al: The TWIST1 oncogene is a direct target of hypoxia-inducible factor-2alpha. Oncogene. 27:1501–1510. 2008. View Article : Google Scholar

125 

Thiery JP: Epithelial-mesenchymal transitions in tumour progression. Nat Rev Cancer. 2:442–454. 2002. View Article : Google Scholar : PubMed/NCBI

126 

Sahlgren C, Gustafsson MV, Jin S, Poellinger L and Lendahl U: Notch signaling mediates hypoxia-induced tumor cell migration and invasion. Proc Natl Acad Sci USA. 105:6392–6397. 2008. View Article : Google Scholar : PubMed/NCBI

127 

Lakatos PL and Lakatos L: Risk for colorectal cancer in ulcerative colitis: Changes, causes and management strategies. World J Gastroenterol. 14:3937–3947. 2008. View Article : Google Scholar : PubMed/NCBI

128 

Munkholm P: Review article: The incidence and prevalence of colorectal cancer in inflammatory bowel disease. Aliment Pharmacol Ther. 18(Suppl 2): 1–5. 2003. View Article : Google Scholar : PubMed/NCBI

129 

Fernández-Majada V, Aguilera C, Villanueva A, et al: Nuclear IKK activity leads to dysregulated notch-dependent gene expression in colorectal cancer. Proc Natl Acad Sci USA. 104:276–281. 2007. View Article : Google Scholar :

130 

Seril DN, Liao J, Yang GY and Yang CS: Oxidative stress and ulcerative colitis-associated carcinogenesis: Studies in humans and animal models. Carcinogenesis. 24:353–362. 2003. View Article : Google Scholar : PubMed/NCBI

131 

Sangha S, Yao M and Wolfe MM: Non-steroidal anti-inflammatory drugs and colorectal cancer prevention. Postgrad Med J. 81:223–227. 2005. View Article : Google Scholar : PubMed/NCBI

132 

Hoffmeister M, Chang-Claude J and Brenner H: Do older adults using NSAIDs have a reduced risk of colorectal cancer. Drugs Aging. 23:513–523. 2006. View Article : Google Scholar

133 

Becker C, Fantini MC, Schramm C, et al: TGF-beta suppresses tumor progression in colon cancer by inhibition of IL-6 trans-signaling. Immunity. 21:491–501. 2004. View Article : Google Scholar : PubMed/NCBI

134 

Greten FR, Arkan MC, Bollrath J, et al: NF-kappaB is a negative regulator of IL-1beta secretion as revealed by genetic and pharmacological inhibition of IKKbeta. Cell. 130:918–931. 2007. View Article : Google Scholar : PubMed/NCBI

135 

Karin M, Cao Y, Greten FR and Li ZW: NF-kappaB in cancer: From innocent bystander to major culprit. Nat Rev Cancer. 2:301–310. 2002. View Article : Google Scholar : PubMed/NCBI

136 

Richmond A: Nf-kappa B, chemokine gene transcription and tumour growth. Nat Rev Immunol. 2:664–674. 2002. View Article : Google Scholar : PubMed/NCBI

137 

Hanahan D and Weinberg RA: The hallmarks of cancer. Cell. 100:57–70. 2000. View Article : Google Scholar : PubMed/NCBI

138 

Schulze-Bergkamen H and Krammer PH: Apoptosis in cancer-implications for therapy. Semin Oncol. 31:90–119. 2004. View Article : Google Scholar : PubMed/NCBI

139 

Kucharczak J, Simmons MJ, Fan Y and Gélinas C: To be, or not to be: NF-kappaB is the answer - role of Rel/NF-kappaB in the regulation of apoptosis. Oncogene. 22:8961–8982. 2003. View Article : Google Scholar : PubMed/NCBI

140 

Luo JL, Kamata H and Karin M: IKK/NF-kappaB signaling: Balancing life and death - a new approach to cancer therapy. J Clin Invest. 115:2625–2632. 2005. View Article : Google Scholar : PubMed/NCBI

141 

Cornejo MG, Boggon TJ and Mercher T: JAK3: A two-faced player in hematological disorders. Int J Biochem Cell Biol. 41:2376–2379. 2009. View Article : Google Scholar : PubMed/NCBI

142 

Lin Q, Lai R, Chirieac LR, et al: Constitutive activation of JAK3/STAT3 in colon carcinoma tumors and cell lines: Inhibition of JAK3/STAT3 signaling induces apoptosis and cell cycle arrest of colon carcinoma cells. Am J Pathol. 167:969–980. 2005. View Article : Google Scholar : PubMed/NCBI

143 

Tsareva SA, Moriggl R, Corvinus FM, Wiederanders B, Schütz A, Kovacic B and Friedrich K: Signal transducer and activator of transcription 3 activation promotes invasive growth of colon carcinomas through matrix metalloproteinase induction. Neoplasia. 9:279–291. 2007. View Article : Google Scholar : PubMed/NCBI

144 

Guttridge DC, Albanese C, Reuther JY, Pestell RG and Baldwin AS Jr: NF-kappaB controls cell growth and differentiation through transcriptional regulation of cyclin D1. Mol Cell Biol. 19:5785–5799. 1999.PubMed/NCBI

145 

Chen C, Edelstein LC and Gélinas C: The Rel/NF-kappaB family directly activates expression of the apoptosis inhibitor Bcl-x(L). Mol Cell Biol. 20:2687–2695. 2000. View Article : Google Scholar : PubMed/NCBI

146 

Baldwin AS: Control of oncogenesis and cancer therapy resistance by the transcription factor NF-kappaB. J Clin Invest. 107:241–246. 2001. View Article : Google Scholar : PubMed/NCBI

147 

Choo MK, Sakurai H, Kim DH and Saiki I: A ginseng saponin metabolite suppresses tumor necrosis factor-α-promoted metastasis by suppressing nuclear factor-κB signaling in murine colon cancer cells. Oncol Rep. 19:595–600. 2008.PubMed/NCBI

148 

Thomasova D, Mulay SR, Bruns H and Anders HJ: p53-independent roles of MDM2 in NF-κB signaling: Implications for cancer therapy, wound healing, and autoimmune diseases. Neoplasia. 14:1097–1101. 2012.

149 

Puvvada SD, Funkhouser WK, Greene K, Deal A, Chu H, Baldwin AS, Tepper JE and O’Neil BH: NF-κB and Bcl-3 activation are prognostic in metastatic colorectal cancer. Oncology. 78:181–188. 2010. View Article : Google Scholar :

150 

Kwon HC, Kim SH, Oh SY, et al: Clinicopathological significance of nuclear factor-kappa B, HIF-1 alpha, and vascular endothelial growth factor expression in stage III colorectal cancer. Cancer Sci. 101:1557–1561. 2010. View Article : Google Scholar : PubMed/NCBI

151 

Wu Y and Zhou BP: TNF-alpha/NF-kappaB/Snail pathway in cancer cell migration and invasion. Br J Cancer. 102:639–644. 2010. View Article : Google Scholar : PubMed/NCBI

152 

Schwitalla S, Ziegler PK, Horst D, et al: Loss of p53 in enterocytes generates an inflammatory microenvironment enabling invasion and lymph node metastasis of carcinogen-induced colorectal tumors. Cancer Cell. 23:93–106. 2013. View Article : Google Scholar : PubMed/NCBI

153 

Terzic J, Grivennikov S, Karin E and Karin M: Inflammation and colon cancer. Gastroenterology. 138:2101–2114. e21052010. View Article : Google Scholar : PubMed/NCBI

154 

Newton IP, Kenneth NS, Appleton PL, Näthke I and Rocha S: Adenomatous polyposis coli and hypoxia-inducible factor-1{alpha} have an antagonistic connection. Mol Biol Cell. 21:3630–3638. 2010. View Article : Google Scholar : PubMed/NCBI

155 

Bienz M and Clevers H: Linking colorectal cancer to Wnt signaling. Cell. 103:311–320. 2000. View Article : Google Scholar : PubMed/NCBI

156 

McCartney BM and Näthke IS: Cell regulation by the Apc protein Apc as master regulator of epithelia. Curr Opin Cell Biol. 20:186–193. 2008. View Article : Google Scholar : PubMed/NCBI

157 

Wen D, Zhang N, Shan B and Wang S: Helicobacter pylori infection may be implicated in the topography and geographic variation of upper gastrointestinal cancers in the Taihang Mountain high-risk region in northern China. Helicobacter. 15:416–421. 2010. View Article : Google Scholar : PubMed/NCBI

158 

Krüger B, Krick S, Dhillon N, et al: Donor Toll-like receptor 4 contributes to ischemia and reperfusion injury following human kidney transplantation. Proc Natl Acad Sci USA. 106:3390–3395. 2009. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

April-2015
Volume 35 Issue 4

Print ISSN: 1107-3756
Online ISSN:1791-244X

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Biddlestone J, Bandarra D and Rocha S: The role of hypoxia in inflammatory disease (Review). Int J Mol Med 35: 859-869, 2015
APA
Biddlestone, J., Bandarra, D., & Rocha, S. (2015). The role of hypoxia in inflammatory disease (Review). International Journal of Molecular Medicine, 35, 859-869. https://doi.org/10.3892/ijmm.2015.2079
MLA
Biddlestone, J., Bandarra, D., Rocha, S."The role of hypoxia in inflammatory disease (Review)". International Journal of Molecular Medicine 35.4 (2015): 859-869.
Chicago
Biddlestone, J., Bandarra, D., Rocha, S."The role of hypoxia in inflammatory disease (Review)". International Journal of Molecular Medicine 35, no. 4 (2015): 859-869. https://doi.org/10.3892/ijmm.2015.2079