Next Article in Journal
Preimplantation Genetic Testing for Aneuploidy (PGT-A) Reveals High Levels of Chromosomal Errors in In Vivo-Derived Pig Embryos, with an Increased Incidence When Produced In Vitro
Next Article in Special Issue
Differential Occupancy and Regulatory Interactions of KDM6A in Bladder Cell Lines
Previous Article in Journal
Metabolic Disruption of Gold Nanospheres, Nanostars and Nanorods in Human Metastatic Prostate Cancer Cells
Previous Article in Special Issue
Pathological Role of HDAC8: Cancer and Beyond
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Transcriptional Regulation during Aberrant Activation of NF-κB Signalling in Cancer

1
School of Biological Sciences (SBS), Nanyang Technological University (NTU), 60 Nanyang Drive, Singapore 637551, Singapore
2
Institute of Molecular and Cell Biology (IMCB), A*STAR, Singapore 138673, Singapore
*
Author to whom correspondence should be addressed.
Cells 2023, 12(5), 788; https://doi.org/10.3390/cells12050788
Submission received: 1 February 2023 / Revised: 16 February 2023 / Accepted: 1 March 2023 / Published: 2 March 2023
(This article belongs to the Special Issue Transcription and Chromatin Dysregulation in Cancer)

Abstract

:
The NF-κB signalling pathway is a major signalling cascade involved in the regulation of inflammation and innate immunity. It is also increasingly recognised as a crucial player in many steps of cancer initiation and progression. The five members of the NF-κB family of transcription factors are activated through two major signalling pathways, the canonical and non-canonical pathways. The canonical NF-κB pathway is prevalently activated in various human malignancies as well as inflammation-related disease conditions. Meanwhile, the significance of non-canonical NF-κB pathway in disease pathogenesis is also increasingly recognized in recent studies. In this review, we discuss the double-edged role of the NF-κB pathway in inflammation and cancer, which depends on the severity and extent of the inflammatory response. We also discuss the intrinsic factors, including selected driver mutations, and extrinsic factors, such as tumour microenvironment and epigenetic modifiers, driving aberrant activation of NF-κB in multiple cancer types. We further provide insights into the importance of the interaction of NF-κB pathway components with various macromolecules to its role in transcriptional regulation in cancer. Finally, we provide a perspective on the potential role of aberrant NF-κB activation in altering the chromatin landscape to support oncogenic development.

1. Introduction

The nuclear transcription factor NF-κB was discovered in 1986 as a Nuclear Factor that binds an immunoglobulin kappa light chain of activated B-cells [1]. NF-κB was subsequently reported to regulate the expression of various important target genes having diverse physiological functions in multiple cell types through its specific DNA binding activity [2,3]. The family of NF-κB transcription factors in mammals comprises five members—RelA (p65), RelB, c-Rel, NF-κB1 (p105/p50) and NF-κB2 (p100/p52) [4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21]. Activation of NF-κB occurs via two major signalling pathways—the canonical and non-canonical pathways that involve distinct regulatory mechanisms and NF-κB members. One of the prominent features of NF-κB transcription factors is their association with the member protein of IκB inhibitor family in the cytoplasm making them unavailable for transcriptional activation in the nucleus. The IκB family typically consists of five members (IκBα, IκBβ, IκBε, IκBζ and BCL3), all sharing similar structures. However, unprocessed p100 and p105 proteins are also categorized as members of the IκB family of proteins due to the presence of typical ankyrin repeats (ANK) in their C-terminal region. An alternative transcript of p105 gene, only reported to be expressed in some murine lymphoid cells, has also been named as one of the members of the IκB family (IκBγ) [22,23,24]. Hence, activation of both the canonical and non-canonical NF-κB pathway involves phosphorylation-dependent degradation of IκB factors by stimulus-response-activated IκB kinases (IKKs). The canonical NF-κB pathway is mediated through the activation of NF-κB essential modifier (NEMO)-dependent IKK (IKKγ), whereas non-canonical NF-κB pathway activation requires a NEMO-independent kinase complex involving IκB kinase α (IKKα) and the NF-κB-inducing kinase (NIK) [25]. Upon activation of the canonical NF-κB pathway, IκB kinases (IKKα, IKKβ and NEMO) phosphorylates inhibitory IκBs and target the latter for proteasomal degradation, resulting in the subsequent nuclear accumulation of NF-κB dimers [26,27,28,29]. In the non-canonical NF-κB pathway, NIK phosphorylates IKKα on Ser 176 position, which in turn phosphorylates p100 subunit, leading to cleavage and ubiquitination mediated degradation of C-terminal half of p100 protein generating active p52 subunit [30]. However, reports also suggest the presence of atypical nuclear-localized IκB proteins, referred to as the BCL3 subfamily (Bcl3, IκBNS, IκBζ and IκBη). These IκBs are reported to show entirely different sub-cellular localization, activation kinetics and functional diversity. They are not only capable of interacting with NF-κB transcription factors inside the nucleus but are also found to get induced and not degraded after NF-κB activation, compared to typical IκB members. In addition, they do not exclusively act as inhibitors of the NF-κB pathway, instead they can regulate the transcriptional activity of NF-κB transcription factors both positively and negatively [25,31,32,33,34]. Nuclear-localized BCL3 act as transcriptional coactivators by removing suppressive p50/p50 homodimers from the promoter of its target genes, in turn allowing binding of activating p50/p65 heterodimer [35]. Bcl3 is also reported to suppress transcription via blocking of the ubiquitination of p50 to stabilize a suppressive NF-κB complex in the nucleus [36]. One interesting finding on the nuclear role of IκB family proteins is its direct binding to NF-κB target sites. Wang et al., showed that Bcl3 forms a complex with p52 homodimer to activate transcription when bound to G/C-rich κB sites in the DNA, whereas the same complex represses transcription when bound to A/T-centric sites in the DNA [37]. In a recent finding, it is reported that atypical IκB Bcl3 enhances the generation of p52 homodimer, subsequently upregulating the expression of target genes involved in proliferation, migration and inflammation [38]. Another BCL3 family member protein, IκBζ, is reported to inhibit transactivation of p65 and its DNA-binding activity in the nucleus [39]. In contrast, IκBη have been shown to be a positive regulator of NF-κB-mediated expression of pro-inflammatory cytokines [40]. Nuclear IκBNS is also shown to interact with several different NF-κB factors in the nucleus but its biological role towards the activity of NF-κB transcription factors is yet to be elucidated [41,42].
In the presence of activating stimuli, the NF-κB-signalling cascade can be induced via either of the pathways depending on the type of stimuli, dimers formed and kinases involved in the post transcriptional modification (PTM) of IκBs and processing of NF-κB factors. In addition to innate and adaptive immune response-dependent activation of the NF-κB pathway, the range of stimuli activating either the canonical or non-canonical pathway varies to a large extent. The canonical pathway of NF-κB is highly inducible and is activated by a diverse range of stimuli, such as radiation, DNA damage, cytokines (TNF-a, IL-1, IL-6), chemokines (MCP-1, IL-8), growth factors, adhesion molecules (ICAM-1, VCAM-1, ELAM), reactive oxygen species (ROS), pattern-recognition receptors (PRRs) and pro-inflammatory receptors such as TNF receptor superfamily (TNFRs) and Toll Like receptor superfamily (TLRs) [43,44,45,46,47,48,49,50,51]. In contrast, the non-canonical NF-κB pathway relies on specific sets of cytokine/receptor molecules for its activation, such as tumour necrosis factor (TNF) receptor superfamily proteins, including BAFF receptor (BAFFR), CD-40, lymphotoxin β receptor (LTβR), Fn14 and receptor activator of nuclear factor kappa-B (RANK) [52,53,54,55], all of which signal through a MAP3K member kinase (MAP3K14) called NF-κB-inducing kinase (NIK), making it a master regulator of the non-canonical NF-κB pathway [56,57,58,59].
Once activated, each subunit of NF-κB signalling cascade, p65 (RelA), RelB, c-Rel, p105/p50 (NF-κB1) and p100/52 (NF-κB2) associate with each other to form distinct transcriptionally active homo/heterodimers [60]. Though they all possess a conserved 300-amino-acid-long amino-terminal Rel homology domain (RHD) that is important for dimerization, DNA binding and interaction with IκBs, as well as nuclear translocation, the role of transactivation is characterised to specific members. RelA (p65), RelB and c-Rel contain the carboxy-terminal transactivation domains (TAD), which form transcriptionally active heterodimers only with p50 and p52 subunits, in turn assisting in DNA-binding activity and activated target gene expression. [61,62]. Reports also suggest formation of homodimers within Rel proteins as well as p50 and p52 subunits [63,64,65,66]. Gourisankar and his group have solved the crystal structures of homodimers of several NF-κB pathway factors such as p65 (RelA) homodimer in complex with a DNA target (2.4 Å resolution) and p50 homodimer bound to a palindromic κB site (2.3 Å resolution) [12,65]. Interestingly, p50/p50 homodimer has been described to exert inhibitory effects on NF-ĸB regulated gene expression [67,68]. c-Rel homodimer is also reported to have the ability to bind IκBα, which in turn inhibits its DNA binding but not cytoplasmic retention [69,70]. A recent report also showed atypical IĸB protein, Bcl3-mediated enhanced generation of p52 homodimer, in turn enhancing transcription of genes involved in cancer-associated biological processes [38]. There are also many other combinations of dimers reported but the most prominent dimers are the RelA-p50 and RelB-p52 dimers, which are activated by the respective canonical and non-canonical pathways [71,72]. The complexity of the NF-κB-signalling mechanisms is further illustrated by the specificity of NF-κB dimers in the transcriptional activation of different target genes.
In addition, NF-κB member proteins also undergo various post-translational modifications (PTMs), like phosphorylation and acetylation, regulating their interaction and crosstalk with components of other signalling pathways. As mentioned earlier, the phosphorylation status of the IκBs determines the activation state of the NF-κB pathway. The availability and activity of NIK, one of the major activating components of the non-canonical pathway also depends on the PTM state of members of its degradation complex containing TRAF3/TRAF2/cIAP1/cIAP2 proteins, which keeps the level of NIK low under constitutive conditions. Upon activation, the degradation complex is recruited to the active receptor complex. This leads to the degradation of cIAP1-cIAP2, thus allowing NIK to dissociate from the complex and subsequently activate the non-canonical NF-κB pathway [73,74,75,76,77,78,79]. However, in an interesting finding in both normal B cells and B cell-derived tumors, it has been shown that CD40 or BAFF receptor activation results in the complete degradation of TRAF3 and partial degradation of TRAF2 but not cIAP1-cIAP2. These findings demonstrate a ubiquitination cascade in which TRAF2 ubiquitinates and activates cIAP1-cIAP2, which then ubiquitinates TRAF3, leading to its degradation and enhanced NIK stabilization as well as processing of NF-κB2/p100 [80]. Low levels of TRAF proteins lead to the higher accumulation of NIK, which then phosphorylates p100 and IKKα, thereby activating the kinase activity important for multiple site phosphorylation of p100 at its C-terminal. Phosphorylated p100 is ubiquitinated by β-TrCP, leading to cleavage-dependent ubiquitination-mediated degradation of C-terminal part of p100, generating active p52 subunit [56,81]. The processing of p100 is important in context to various steps of regulation in the activation of NF-κB pathway. Unprocessed p100 binds RelA, RelB or c-Rel subunit via its C-terminal ankyrin repeats, further inhibiting the activity of Rel subunits [82,83]. Hence, in the context of activation of the non-canonical NF-κB pathway, stabilization of NIK and processing of p100 acts as one of the major steps involving multiple PTMs of its regulator molecules [58].
Specific stimuli-dependent activation of the non-canonical NF-κB pathway is important in regulating various important biological functions such as lymphoid organogenesis, B-cell survival and maturation, dendritic cell activation and bone metabolism [84,85,86,87,88,89,90]. In spite of being tightly regulated by various activators and inhibitory factors, the aberrant activation of the NF-κB pathway has been observed in many lymphoid malignancies. Besides its role in immune regulation, NF-κB members have been documented to regulate transcriptional activities that promote the malignant transformation and survival of cancer cells. Several studies demonstrate the presence of promiscuous mutations responsible for inhibiting TRAF2, TRAF3 and cIAP1/2 complex or enhancing the expression/stability of NIK and other receptor molecules like CD-40 and LTβR. Such mutations are associated with the abnormal activation of the non-canonical NF-κB pathway [85,91,92,93,94,95]. Additionally, recent studies suggest the interdependency of NF-κB-driven expression of target genes with epigenetic changes in the genome. In this review, we will discuss the activation and regulation of NF-κB signalling in inflammation and cancer in context to its interaction with transcription factors (TFs), kinases, epigenetic modifiers and non-coding RNAs. We focus on discussing the interdependent role of NF-κB-signalling components with transcription factors and chromatin modifiers in the aberrant activation of the NF-κB pathway, as well as in the active transcriptional activation of its target genes (summarized in Figure 1).

2. NF-κB: One of the Key Factors Linking Inflammation in Cancer

Even before Vichow’s hypothesis on the origin of cancer from the site of inflammation, several inflammation-associated viral and bacterial infections (Hepatitis B, Helicobacter pyroli) were found to be associated with increased risk of malignancies of the liver, colon and stomach [96,97,98,99]. Additionally, statistical reports estimate that inflammatory viral infections contribute to >15% of all cancers [100,101]. It has been postulated that cancer cells can hijack the normal inflammatory mechanism to boost their growth and survival. In general, the normal function of immune cells is to trigger the innate and adaptive immune response to differentiate between self/non-self and destroy/engulf the foreign invaders. However, in the tumour microenvironment, cancer cells can alter the protective functions of immune cells and convert them to act as tumour-promoting cells. They are reprogrammed to secrete pro-survival inflammatory cytokines, allowing better proliferation, survival, migration, invasion and inhibition of apoptosis in cancer cells. Hence, tumour-promoting inflammation acts as one of the major hallmarks of cancer. But what remains unanswered are the regulatory molecules that link inflammation to cancer progression. One of the major pathways reported to be involved in manipulating the immune response machinery in the tumour microenvironment is the NF-κB pathway.
Several studies have converged on the role of the NF-κB pathway as one of the critical missing links between inflammation and cancer. The first evidence comes from various studies reporting extensive sequence similarity between c-Rel and viral oncoprotein v-Rel in their N-terminal domain, a region referred to as the Rel Homology Region (RHR), and identification of oncogene Bcl3 as a member of the IκB family [102,103]. In addition, many cancer cell types show elevated levels of NF-κB expression and activation. Endogenous activation of NF-κB is reported in Hs294T melanoma cells due to altered equilibrium between IκBα degradation and resynthesis, leading to overall decrease in the level of IκBα expression [104]. Moreover, human colorectal cancer (CRC) epithelial cells have been observed to express enhanced NF-κB and IκBα, which is accompanied by the increased expression of cox-2 gene [105]. Similar selective activation of the NF-κB pathway is also reported in breast cancer. The RelA subunit of NF-κB is reported to be activated in breast cancer cell lines, whereas breast tumours are shown to exhibit an absence or low level of nuclear RelA, in contrast to activated c-Rel, NF-κB1 and NF-κB2 along with bcl2 expression, as compared to nontumorigenic adjacent tissue [106]. Most interestingly, the NF-κB family of transcription factors have been shown to contribute to the function and maintenance of tumour-initiating cells (TICs) in breast cancer. Experimental data indicates the activation of both the canonical and non-canonical NF-κB pathway to be important in the function of TICs by stimulating epithelial-to-mesenchymal transition (EMT) and upregulating the expression of the inflammatory cytokines IL-1β and IL-6 [107]. In a recent finding, Monica et al. reported the involvement of the NF-κB pathway towards resistance to endocrine and chemotherapies in breast cancer [108]. Furthermore, BRCA1 signalling, which is one of the prominent pathways activated in various cancer types including breast cancer and ovarian cancer, additionally possesses the capability to induce the NF-κB pathway [109].
Additionally, in lung carcinoma, the enhanced expression of IKKβ and NF-κB is reported as an important factor for tumour initiation and progression [110]. Another NF-κB activity dependent type of cancer is melanoma. Studies using mice models have shown that initiation of such tumours is HRas-mediated and involves the regulation by IKKβ in the activation of NF-κB [111]. In the cell line model of Diffuse large B-cell lymphoma (DLBCL), constitutive activity of IKK and high NF-κB DNA-binding activity is reported in the ABC-DLBCL subtype but not in the GC-DLBCL subtype [112]. Recently, Eluard et al. reported the presence of a new subset of Diffuse large B-cell lymphoma (DLBCL) patients showing enhanced RelB activation with aberrant gene expression and mutation profiles [113]. Based on these studies, it can be postulated that abnormal activation of the NF-κB pathway in various cancer types is critical for the survival of transformed cells, particularly in the suppression of apoptosis and senescence. Apart from endogenous elevated levels of NF-κB expression and activation in cancers (summarized in Table 1), there are also reports on aberrant activation of NF-κB pathway in cancer cells. Hence, there appears to be various extrinsic and intrinsic factors regulating the malignancy-associated enhanced activation of the NF-κB pathway in cancers.

3. Factors Contributing to the Hyperactivation of NF-κB Pathway in Cancers

3.1. Cancer Associated Immune Signalling Molecules

In addition to the role of elevated NF-κB activity in the survival of transformed cells, NF-κB is found to be activated in cancer stem cells (CSCs). In CSCs, it promotes the release of pro-inflammatory cytokines that exert anti-apoptotic and pro-proliferative activities [129]. One of the prominent factors involved in the activation of the canonical NF-κB pathway in both solid and hematologic malignancies is the tumour microenvironment (TME). A large number of immune cells (macrophages, dendritic cells, neutrophils, mast cells, T cells and B cells) are recruited to the TME, leading to the enhanced production of cytokines, growth and angiogenic factors and proteases that degrade the extracellular matrix to support cancer development and progression [130]. In solid tumours, the sustained activation of the NF-κB pathway is predominantly achieved through the continuous release of cytokines by tumour-associated macrophages (TAM) in the TME. One of the predominant properties of TAM is the ability to switch from M1- to M2-phenotype with enhanced release of anti-inflammatory cytokines [131,132], suggesting a crosstalk between cancer cells and neighboring macrophages. Interestingly, IKKβ and NF-κB are also reported to assist in the polarization of macrophages towards the M2 type, which fosters and protects the tumour cells instead of attacking them [133,134]. Hence, this permits malignant cells to bypass tumour immunosurveillance activity via NF-κB-mediated polarization of macrophages from the M1 to M2 phenotype.
Activation of either of the canonical or non-canonical pathways in both solid and hematologic malignancies also depends on different sets of inflammation-associated cytokines and receptors activated in tumour cells. Induction of the canonical NF-κB pathway is initiated by pattern recognition receptors and diverse tumour-promoting cytokines, such as TNF, IL-1, and IL-17 [135]. On the contrary, activation of the non-canonical NF-κB pathway is triggered by signalling via a specific subset of TNFR superfamily members such as B-cell-activating factor belonging to TNF family receptor (BAFFR) [52,55], CD-40 [53], lymphotoxin β-receptor (LTβR) [54], receptor activator for nuclear factor κB (RANKL) [136], TNFR2 [137,138], Fn14 [139], etc.
In addition to the involvement of immune regulatory molecules in activation of both the canonical and non-canonical NF-κB pathways towards cancer-promoting mechanisms rather than their classical immunosurveillance roles, the important observation is their expression in non-immune cells. Enhanced expression of CD-40 is reported in many non-immune cells, such as the intestinal epithelial cells (IECs) of patients with colon cancer. This, in turn, leads to the aberrant activation of non–canonical NF-κB pathway, suggesting the important link between immunosurveillance and tumorigenicity [140,141,142]. LTβR is expressed in lymphoid stromal and epithelial cells. BAFFR is predominantly expressed in B cells, whereas RANK, which is best known for its role in osteoclastogenesis, is also reported to be highly expressed in various cancer types like breast and prostate cancer cells, mediating the migration and skeletal metastasis of cancer cells [143,144].

3.2. Intrinsic Mutations and Post Translational Modifications (PTMs)

In addition to the receptor dependent aberrant activation of NF-κB pathways in cancer, activating mutations in other signalling components of the non-canonical NF-κB pathway have been documented particularly in lymphoid malignancies [91]. Such activation is driven by the presence of selected mutations inactivating the genes encoding negative regulators of the pathway (TRAF2, TRAF3, TNFAIP3, BIRC3, MAP3K14, CYLD, cIAP1/cIAP2) and activating the regulator molecules (NF-κB1, NF-κB2, CD40, LTβR, TACI, and NIK) in various cancer types like multiple myeloma (MM), splenic marginal zone lymphoma (SMZL), MALT lymphoma and B-cell lymphoma [15,93,145,146,147,148,149] (Table 2). Mutations leading to constitutive activation of the kinase NIK in multiple myeloma have been found in NIK itself, that disrupts its binding with TRAF3, in turn causing dissociation of NIK from the inhibitory complex having TRAF2 and the ubiquitin ligases cIAP1 and cIAP2. This, in turn, results in NIK stabilization, leading to aberrant activation of the non-canonical NF-κB pathway [93,119]. The genetic selection of these driver mutations by cancer cells highlights the critical importance of the NF-κB pathway towards cancer progression and enhanced malignancy. In the case of multiple myeloma, mutations are also reported in many other signalling subunits of non-canonical NF-κB pathway—NF-κB2, TRAF2, TRAF3, BTRC encoding β-TrCP, which alters the inhibitory degradative pathway of NIK kinase by TRAF2/TRAF3 complex, leading to malignancy-associated activation of non-canonical NF-κB signalling [150]. Overexpression of NIK due to t(17;22) chromosomal translocation is also associated with the occurrence of multiple myeloma [93]. Oncogenic mutations in the TP53 protein are reported to be associated with higher RelA expression, in turn activating the canonical NF-κB pathway in human B-cell lymphomas such as Hodgkin lymphoma and, to a lesser extent in T-cell lymphoma cell lines as well [151,152]. Chromosomal translocation t(10;14)(q24;q32) of the NF-κB2 gene is observed to be associated with a variety of hematological malignancies, such as MALT lymphomas [153]. The translocation moves the IgG promoter to a region upstream of the bcl-10 gene, resulting in expression of a truncated bcl-10 protein, leading to activation of NF-κB [153]. Another reported translocation is t(11;18), which results in the generation of a chimeric protein, AP12-MALT1, which leads to NF-κB activation in B-cell lymphomas [154]. The modified NF-κB2 gene codes for the protein that lacks the ankyrin regulatory domain but still binds the kappa B sequence in vitro. Such rearrangement of NF-κB2 has been reported in both B-cell and T-cell lymphoma patients suggesting that translocation dependent truncation of the ankyrin domain may be a common mechanism in the abnormal activation of NFKB2 gene and its relevant role in lymphomagenesis [15,148,153,155,156,157]. Another genomic rearrangement event reported in DLBCL is on chromosome 10q24, which results in increased NFKB2 mRNA expression, causing constitutive expression of NF-κB2 [15,158]. Chromosomal translocation of the c-Rel gene to chromosome 2p 13–15 causing its enhanced amplification has been reported in DLBCLs with a large cell component, constituting approximately 50% of B-cell non-Hodgkin’s lymphomas [149,159,160,161]. This chromosomal aberration of c-Rel has also been found in primary mediastinal (thymic) B-cell lymphomas and follicular large cell lymphomas, and is reported to be associated with extra nodal presentation [120]. Another member of the NF-κB family, RelA, is mapped to be translocated to 11q13, a site where a number of genes involved in neoplastic development have already been mapped, suggesting a link between chromosomal translocation and the tumour-inducing role of RelA [162]. Activating mutations (translocation t(14;19)(q32;q13.1)) in another member of the NF-κB pathway, Bcl3, which is a proto-oncogene, have also been observed in B-cell leukaemia. The chromosomal translocation of Bcl3 results in its enhanced expression in leukemic cells as compared to normal blood cells [163]. In addition to biallelic deletion and chromosomal translocation, several other mutations, including missense mutation, frameshift mutation and in frame deletion are also reported to inactivate the TRAF3 gene, in turn inducing the activation of NF-κB pathway [91].
Besides the reported genomic abnormalities, there are several other factors that can influence the transcriptional activity of NF-κB pathway. Reports suggest that crosstalk with certain activating and inhibitory kinases such as Glycogen Synthase Kinase (GSK-3β), p38 and PI3K can either modulate the transcriptional activity of NF-κB or its upstream signalling pathways [164,165,166]. Kinases are reported to modulate the activity of NF-κB in glioma cell lines and pancreatic cancer cells through post-translational modification (PTM) of the NF-κB subunits (p65/p50) [166,167,168,169]. Further studies indicate that GSK-3β has no role in the nuclear accumulation of NF-κB, but instead alters the DNA-binding activity of NF-κB subunits by inducing hypermethylation of the target DNA [167,168,169,170,171]. Other kinases documented to regulate the NF-κB pathway include the Jun- N-terminal kinase (JNK) and p38 [172]. Though both the kinases can be induced by the same stimuli (TNFα) that activate NF-κB pathway, they have been found to display differential functions on NF-κB activity. p38 acts as a co-factor to modulate the transactivation machinery of NF-κB to regulate TNF-induced IL-6 gene expression, whereas a counteracting relationship occurs between JNK and NF-κB. NF-κB complexes downregulate the c-Jun amino-terminal kinase (JNK) cascade via upregulation of gadd45β/myd118 gene expression. Gadd45β, in turn, targets MKK7/JNKK2, a specific and essential activator of JNK. Mechanistically, binding of gadd45β with MKK7 blocks the catalytic activity of the latter, causing inhibition of the JNK pathway [173,174,175,176]. Hence, the aberrant activation of NF-κB pathway depends on multiple factors including cell type, micro-environment, PTMs, enzymatic activity of regulatory molecules and chromosomal abnormalities.
Table 2. Summarized table of various mutations/chromosomal alterations in components of NF-κB pathway affecting its expression and activity in different cancer types.
Table 2. Summarized table of various mutations/chromosomal alterations in components of NF-κB pathway affecting its expression and activity in different cancer types.
Factors/Regulators of NF-κB PathwayType of Mutation/Chromosomal TranslocationCancer TypeResponseReferences
TRAF3Bi-allelic deletion at 14q32MMIncreased p52/p100 ratios[91,177]
TRAF2Bi-allelic deletion at 9q34MMIncreased p52/p100 ratios[91,177]
CYLDBi-allelic deletion at 16q12MMIncreased p52/p100 ratios[91,177]
cIAP1/cIAP2Bi-allelic deletion at 11q22MMIncreased p52/p100 ratios[91,177]
NIKt(17;22) translocation, IgH translocation or amplificationMMOverexpression of NIK[93]
LTBRamplification of the entire 12p chromosome armMMActivatory[91]
NF-κB2t(10;14)(q24;q32)
t(10q24)
MALT Lymphoma,
DLBCL
Activatory,
Enhanced expression of NF-κB2 gene and protein
[15,91,153]
AP12-MALT1t(11;18)(q21;q21)B-cell Lymphoma, MALT lymphomaActivatory[154,155]
c-Relt(2p 13-15)DLBCL, B-cell lymphoma, Follicular large cell lymphomaEnhanced amplification of c-Rel gene[120,121,149,159,161]
RelA(11q13) site with t(11;14)(q13;q32)NHL, Diffuse large cell lymphoma, Squamous carcinoma of head and neck, Breast cancer Activatory[159,162]
Bcl3t(14;19)(q32;q13.1)B-cell leukaemiaActivatory[163]

3.3. Epigenetic Modification in the Component(s) of NF-κB Pathway

Reports suggest the dependency of NF-κB components on various epigenetic factors for its activation in cancer cells. Reduced expression of histone methyltransferase EZH2 stimulates the expression of TRAF2/5 via the de-repression of their expression due to H3K27 hypermethylation by EZH2. The elevated TRAF2/5 expression, in turn, enhances TNFα-induced activation of NF-κB signalling, leading to an uncontrolled inflammatory reaction which ultimately contributes to tumorigenesis [178]. The enhanced activation and expression of NF-κB-signalling component proteins in various cancer types also depends on the epigenetically modified state of its own component genes and its target genes. Triple negative breast cancer cells display a high level of NF-κB activation due to the enhanced expression of NIK that is caused by the epigenetic alteration (histone H3 acetylation) of the NIK gene [179]. Hence, these studies suggest the plausible de-regulation of the NF-κB pathway due to epigenetic alterations.

4. Double Edged Role of NF-κB from Immunosurveillance to Pro-Tumorigenic Functions

As discussed, the aberrant activation of the NF-κB pathway in cancer is a multifactorial event. Depending on the prevalent tumour microenvironment, malignancy-promoting mutations in the components of the NF-κB-signalling cascade, and the inflammatory molecules released by the tumour immune cells, the biological importance of the NF-κB pathway is diverted from the immunosurveillance mechanism towards tumour-promoting functions.
NF-κB signalling has been shown to activate the expression of various inflammatory mediators, such as IL1β, TNF and IL6, which promote cancer development [180,181]. However, the question remains as—what factor(s) drives the variation in inflammatory response by the NF-κB pathway from a protective role towards a tumour-promoting role. The answer to this oncogenic shift is related to the severity of inflammation response which mostly occurs during chronic inflammatory conditions. During acute inflammatory conditions, NF-κB activation acts as a tumour immunosurveillance mechanism to assist in the targeting and elimination of transformed cells. For example, protein kinase D1-mediated activation of NF-κB signalling can induce the expression of antioxidant proteins such as MnSOD and anti-apoptotic proteins including A20 and cIAPs to prevent the accumulation of pro-tumorigenic ROS that can cause oncogenic mutations [182,183,184,185,186]. NF-κB-mediated inhibition of ROS accumulation can also repress the activity of pro-tumorigenic transcription factors such as STAT3 and AP1 [185]. In contrast, under chronic inflammatory conditions, the continuous presence of NF-κB stimuli seem to outperform the inhibitory role of the negative NF-κB regulators, leading to constitutive activation of NF-κB signalling. Such constitutive activity of NF-κB can exert pro-tumorigenic effects ranging from cell proliferation and cell survival to malignant cell invasion and metastasis. Many cancers arise from sites of chronic infection or inflammation due to elevated ROS production by neutrophils in response to invading pathogens. This innate immune response in turn causes DNA damage and genetic mutations, thereby triggering tumour initiation [186,187].
Though both the canonical and non-canonical NF-κB pathways are reported to be activated in various invasive and malignant cancers, the functional mechanism for downstream substrates involved in activation of the non-canonical NF-κB pathway are well characterized compared to the canonical pathway. The invasive nature of Glioblastoma Multiforme (GBM) cells has been reported to be associated with high RelB expression [116,117]. Work on mouse tumour xenograft models also showed activation of the non-canonical NF-κB pathway leading to regulation of the expression of its own regulator protein NIK, which, in turn, is reported to induce dramatic cell shape changes, increase tumour cell invasion and promote aggressive orthotopic tumour growth [123]. Point mutations at the promoter region of the telomerase reverse transcriptase (TERT) gene is one of the most frequent non-coding mutations in cancer. TERT promoter mutations (TPMs) are cancer type-specific and among the first few mutations reported in melanomas, glioblastomas and hepatocellular carcinomas [188,189,190,191]. In an interesting finding, non-canonical NF-κB signalling is reported to drive the expression of the TERT gene carrying −146 C > T mutation in its promoter region, causing telomerase reactivation, which is otherwise not activated via binding of ETS transcription factor [192,193]. This data specifically highlights a novel role of the non-canonical NF-κB pathway in the reactivation of telomerase in cancers. Hence, the level of inflammatory response and genetic changes in the cancer cells can act as some of the major factor(s) deciding the difference between acute inflammatory response versus aberrant activator response of the NF-κB-signalling pathway in cancer.

5. Aberrant NF-κB Activation Driven Expression of Tumour Promoting Genes

Apart from activating the expression of its immune response target genes, aberrantly activated NF-κB signalling in cancer cells contribute to cancer progression by acting as a transcriptional activator of various other pro-tumorigenic genes involved in cell proliferation, inhibition of apoptosis, invasion, metastasis and angiogenesis.
In-depth studies also show that NF-κB controlled genes regulating oncogenic properties are significantly different. NF-κB-dependent cancer-relevant genes mostly encode for cytokines, cell cycle genes like cyclin D1, matrix metalloproteinases (MMPs) and anti-apoptotic proteins. Numerous NF-κB target genes such as cIAP1/2, TRAF1/2, Bcl-xL, XIAP, MnSOD and IEX-1L confer antiapoptotic properties [106,194,195,196]. Specifically, the NF-κB target gene cIAP1/2 functions as an inhibitory factor of cancer cell apoptosis through directly binding and suppressing the effector caspases [197,198]. NF-κB signalling controls the epithelial to mesenchymal transition and metastasis, often via upregulation of matrix metalloproteinases (MMPs) [199]. In breast cancer, NF-κB is also reported to induce the expression of EMT-related genes such as Twist, intercellular adhesion molecule-1 (ICAM-1), endothelial leukocyte adhesion molecule 1 (ELAM-1), vascular cell adhesion molecule 1 (VCAM-1), MMPs and serine protease urokinase-type plasminogen activator (uPA), along with the expression of one of the major tumour promoting genes Bcl2 [200,201]. Interestingly, one study revealed a role for NIK in the phosphorylation, enzymatic activity and pseudopodal localization of membrane type 1 MMP in highly invasive tumours like glioblastoma that is distinct from its established kinase function in the non-canonical NF-κB pathway [202].
NF-κB signalling also contributes to tumour progression and invasion by controlling pro-angiogenic genes such as vascular endothelial growth factor (VEGF) and its receptors, macrophage inflammatory protein-1 (MCP-1) and CXC-chemokine ligand 8, also known as IL-8 (CXCL8) [203,204,205,206,207]. Activated NF-κB signalling in cancer transactivates the expression of cyclin D1 and c-myc that promote cancer cell proliferation [208,209]. Angiogenesis, the phenomenon of new blood vessel formation is one of the hallmark phenotypes of cancer cells. Tumour angiogenesis is dependent on proinflammatory cytokines, chemokines and growth factors such as MCP-1, IL-8, TNF-α and VEGF, secreted by tumour-associated macrophages (TAMs) via the activated NF-κB pathway. Furthermore, the recruitment of bone marrow-derived cells (BMDCs) to tumours for vasculogenesis is essential for tumour angiogenesis, which is found to involve NF-κB-mediated enhanced expression of IL-8 and angiogenin [210,211]. Subsequently, the expression and activation level of different NF-κB subunits can induce varying severity in different cancer types. In the case of ER-positive breast carcinoma, higher expression of RelB is associated with decreased relapse-free survival (RFS) and overall survival (OS) rate, whereas in other tumours, such as lung carcinoma, enhanced expression of NIK and RelB is associated with enhanced metastasis and shorter OS. Poor RFS outcome is reported to be associated with higher expression of non-canonical NF-κB target gene myoglobin [212,213,214]. Elevated RelB activity reported in a new subset of DLBCL patients is found to confer resistance to DNA damage-induced apoptosis along with increased cIAP2 expression [113]. In a more recent finding, sustained activation of the non-canonical NF-κB signalling is also shown to drive doxorubicin resistance in DLBCL via enhanced glycolysis [215]. Hence, these studies indicate the existence of a high degree of NF-κB dysregulation in cancer.

6. Different Modes of Deregulated NF-κB Signalling in Cancer

While we discussed the multifaceted roles of the NF-κB pathway linking inflammation and cancer, it is also important to understand the interacting map of the components of this pathway with other macromolecules, which, in turn, regulate the transcription of pro-oncogenic transcripts (Figure 2).

6.1. Interaction with Transcription Factors

While NF-κB regulates the expression and activity of various regulatory factors, its own activity can also be regulated via direct association with several other transcription factors. The most prominent ones are proto-oncogenic transcription factors such as STAT3, p53, AP1 and ETS-related genes ERG, implicating their plausible cooperative function with NF-κB factors in inflammation and cancer [216,217,218]. Hence, depending on the promoter sequence and structure of the target genes, the functional link between NF-κB and other transcription factors might vary. One of the well-characterized factors known to co-associate with NF-κB is the STAT family members. NF-κB, in association with STAT3, regulates the expression of various cell cycle genes, anti-apoptotic genes and genes encoding cytokines and chemokines [219]. Studies suggest that the direct interaction of RelA and NF-κB1 members with STAT3 facilitates both the recruitment of NF-κB and STAT3 onto each other’s promoter sites [220,221,222]. In another context of regulation, STAT3 modifies the RelA subunit by recruiting acetyltransferase p300, resulting in the acetylation-dependent retention of NF-κB in the nucleus [223]. Such regulation leads to the enhanced activity of NF-κB (a tumour-promoting phenomenon) and hence, chronic stimulation of cytokines in the tumour microenvironment. Cross talk of NF-κB with transcription factor p53 also occurs [221]. Enhanced secretion of the pro-inflammatory cytokine TNFα triggers the formation of an active complex containing nuclear RelA and p53 on κB binding motifs, suggesting the importance of p53 in NF-κB-mediated gene expression induced by canonical stimuli [224,225]. In addition, some reports suggest that the RelA subunit and transcription factor p53 can regulate their respective transcriptional activities. p53 has been shown to inhibit NF-κB transcriptional activity, while the RelA subunit can also inhibit p53-dependent transactivation of target genes [221]. This constitutive activation of NF-κB, evoked by a p53 hot-spot mutant protein frequently found in tumours, provides an explanation for the fact that p53 mutations arise more than p53 deletions in tumours of various origin [222,226]. More recently, another transcription factor, the ETS family member ERG, has been identified to cross talk with NF-κB. As reported by various groups, the functional role of ERG is validated in various leukemia, Ewing sarcoma and prostate cancer [227,228,229,230]. Interestingly, NF-κB activation is elevated in ERG fusion-positive prostate cancer patients and cancer cell lines [231]. ERG is also reported to regulate expression of the NF-κB target gene, ICAM-1 in endothelial cells [232,233]. Another interesting study also revealed the cooperative function of p52 with transcription factor ETS1 in the reactivation of telomerase in cancers via a hotspot −146 C > T TERT promoter mutation [192]. On a similar line, a recent finding has shown the involvement of the non-canonical NF-κB pathway in altering the genomic binding landscape of transcription factor ETS1 that supports glioma progression [234]. Hence, a cross talk is predicted between NF-κB and other TFs at the level of activation and transcriptional regulation of NF-κB target genes, which requires further studies for in depth understanding of the mechanisms involved.

6.2. Effect of Pro Tumorigenic Non-Coding RNAs

Considering the challenge with highly evolving cancer cells which are resistant to many available therapies either through selected genetic mutations or positive adaptation to the cancer microenvironment, it is critical to understand new alternative modes of regulations adopted by cancer cells. In recent times, one such regulatory molecule showing relevance in context to its crosstalk with the components of NF-κB pathway is non-coding RNAs (ncRNAs). Altered regulation at the level of epigenome mediated by non-coding RNAs (microRNAs—miRNAs and long noncoding RNAs—lncRNAs) has been found to be a prevailing factor impacting various types of malignancies. Several miRNAs are transcriptional targets of NF-κB, such as miR-9, miR-21, miR-143, miR-146 and miR-224, which, in turn, act as a feedback mechanism for modulating the activity of NF-κB [235,236,237,238,239,240,241]. Out of these, miR-21 and miR-143 are reported to be involved in regulating the malignant phenotypes like invasion and metastasis in cancer types including breast cancer and HCC [238,239]. On the other hand, NF-κB can also induce the expression of proteins important for the transcriptional regulation of miRNAs. One such example is the NF-κB driven expression of Lin28 protein, which inhibits the processing and maturation of let-7 miRNAs—a family of tumour suppressor miRNAs whose expression is downregulated in many cancer types. Let7 miRNA also targets IL6. Thus, Lin28-mediated downregulation of Let7 miRNA leads to the higher expression of IL6 and further enhances NF-κB signalling in a positive feedback loop mechanism [242].
Subsequently, NF-κB activity is also regulated by the presence of several miRNAs mostly via repressive mechanisms. One such miRNA is miR-502e, which is reported to act as a tumour suppressor factor by altering cell proliferation in hepatoma cell lines and hepatocellular carcinoma by targeting NIK, thereby modulating the activity of non-canonical NF-κB signalling [243]. Many highly expressed long non-coding RNAs (lncRNAs) are also reported to regulate the activity of NF-κB. The lncRNA NKILA, was reported to mask the phosphorylation motifs of IκB, further inhibiting the activation of NF-κB [244]. Along with the aberrant activation of the NF-κB-signalling pathway, the expression of long non-coding RNAs (ncRNAs) is also dysregulated in different types of cancer cells, further regulating the degree of malignancy. The upregulated expression of lncRNA H19 in melanoma cells and Helicobacter pylori-induced expression of H19 in gastric cancer cells have been reported to be associated with enhanced cancer cell invasion and migration via activation of the NF-κB- and PI3K/Akt-signalling pathways [245,246]. Another NF-κB-associated lncRNA reported to be upregulated in cancer cells is lncRNA NEAT1. Its overexpression promotes proliferation, migration and invasion, influences the expression of EMT markers, and activates the NF-κB pathway in HeLa and SiHa cells [247]. H19 and NEAT1 are also reported to be associated with the resistance of cancer cells to chemotherapeutic drugs including bortezomib and dexamethasone respectively [248,249]. Hence, it can be speculated that subunits of NF-κB function in association with ncRNAs to impart their pro-tumorigenic roles along with chemoresistance functions in tumour cells whose mechanism remains elusive and requires further clarification.

7. Role of NF-κB Signalling in Shaping the Cancer Cell Chromatin Landscape

Though the NF-κB family of proteins lack endogenous chromatin modifying enzymatic activity, they can exert changes in the chromatin landscape either by acting as a mediator to recruit and position chromatin modifiers onto target genes in a specific sequence dependent manner or by regulating the expression and activity of those modifiers [63,250]. One noteworthy feature of NF-κB family members is their ability to form multimeric complexes. Apart from forming multimeric complexes with its own family proteins, NF-κB subunits are reported to form complexes with other proteins, which includes chromatin modifiers as well. Upon lymphotoxin treatment, non-canonical NF-κB signalling is activated and RelB/p52 dimer gets associated with the SWI/SNF chromatin remodeling complex via an adapter protein, requiem, to induce the expression of the BLC gene (CXCL13). Such interaction suggests an indirect role of activated NF-κB signalling in the epigenetic regulation of oncogene expression [250]. Additionally, the NF-κB pathway also acts as a key regulator in the enhanced expression of chromatin modifiers and its subunits/interacting proteins, such as Enhancer of Zeste Homologue 2 (EZH2), a histone-lysine N-methyltransferase enzyme involved in the epigenetic modification of histone protein (H3K27), thus conferring the hypermethylation-mediated repressive gene expression of anti-oncogenic genes [251]. In colorectal cancers, NF-κB activation in response to TNFα has been reported to induce the expression of EZH2, leading to the inhibitory promoter hyper-methylation of pro-apoptotic protein kinase cδ binding protein (PRKCDBP) and resultant increased growth of cancer cells [252]. Subunits of the NF-κB pathway can also act in a de-repression mechanism to remove the repressive chromatin marks and complexes. Some inducible gene promoters harbor high levels of the H3K9 dimethyl modification, associated with transcriptional silencing. However, upon stimulation, these marks are removed by the Aof1 histone demethylase, whose recruitment requires initially bound c-Rel dimers within the promoter region [253,254]. The NF-κB pathway is also reported to regulate RNA Polymerase II elongation by changing the chromatin landscape via recruitment of General Control Non-Derepressible 5 (GCN5) acetyltransferase complexes that primarily modify H4K5/K8/K12 lysine residues. The accumulation of acetylated H4 histone proteins leads to the association with BRD4, which then positively regulates transcription by recruiting the elongation factor P-TEFb [255]. Hence, the ability of the components of the NF-κB pathway to alter the chromatin landscape is not only limited to its signature DNA binding property but also extended to the recruitment of various chromatin modifiers assisting in transcriptional regulation.

8. Concluding Remarks

Since the discovery of NF-κB nearly four decades ago, the multi-faceted roles of NF-κB members and their new transcription-binding partners in cancer have been gaining more clinical relevance in recent years. Although inflammation was previously implicated to promote the malignancy of human cancers, the causal mechanisms underscoring the link between inflammation and cancer have not been adequately characterized. Recent studies showing the aberrant activation of the NF-κB pathway in various cancer types and the regulation of NF-κB members in various tumorigenic events support the role of NF-κB as a hub linking inflammation and cancer. Although the occurrence of activating mutations in the NF-κB pathway is predominantly observed in hematological malignancies, the activation of NF-κB in solid tumours is also not negligible. The functional shift of the NF-κB pathway from inflammation to oncogenesis is mostly driven by the onset of chronic inflammatory conditions. NF-κB members can exert pro-oncogenic functions during cancer development through the activation of target gene transcription by their heterodimers. In addition, NF-κB components have also been demonstrated to interact with other factors, including transcription factors, kinases, epigenetic modifiers and other biological molecules like ROS and ncRNAs, to drive multiple oncogenic activities. Despite substantial progress in the understanding of various aspects of NF-κB signalling in cancer, the approaches for the targeted inhibition of specific components in the signalling pathway are limited due to various challenges. These challenges arise from the complex nature of its activity in different cancer types. Recent genomics studies have revealed the active selection of a wide range of driver mutations in cancer cells, some of which are important to facilitate the activation of the NF-κB pathway. In addition, epigenetic alterations have been documented to contribute to the aberrant activation of the NF-κB pathway. Conversely, the activated NF-κB pathway is also reported to confer changes in the chromatin landscape of cancer cells towards enhanced malignant phenotypes. Hence, these findings can potentially pave new ways for the development of precision medicine to improve the efficiency of existing cancer therapies and overcome the phenomenon of multidrug resistance in most of the cancer types.

Author Contributions

Conceptualization, K.D. and Y.L.; writing—original draft preparation, K.D.; writing—review and editing, K.D. and Y.L.; supervision, Y.L.; project administration, Y.L.; funding acquisition, Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work is funded by the National Research Foundation (NRF) Singapore, under its Singapore NRF Fellowship (NRF-NRFF2018-04).

Acknowledgments

We acknowledge the financial support provided by the National Research Foundation (NRF) Singapore, under its Singapore NRF Fellowship (NRF-NRFF2018-04) to cover the funding for this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sen, R.; Baltimore, D. Inducibility of κ immunoglobulin enhancer-binding protein NF-κB by a posttranslational mechanism. Cell 1986, 47, 921–928. [Google Scholar] [CrossRef] [PubMed]
  2. May, M.J.; Ghosh, S. Signal transduction through NF-κB. Immunol. Today 1998, 19, 80–88. [Google Scholar] [CrossRef] [PubMed]
  3. Ghosh, S.; May, M.J.; Kopp, E.B. NF-κB and Rel proteins: Evolutionarily conserved mediators of immune responses. Annu. Rev. Immunol. 1998, 16, 225–260. [Google Scholar] [CrossRef] [PubMed]
  4. Ghosh, S.; Karin, M. Missing pieces in the NF-κB puzzle. Cell 2002, 109, S81–S96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Lin, L.; DeMartino, G.N.; Greene, W.C. Cotranslational biogenesis of NF-κB p50 by the 26S proteasome. Cell 1998, 92, 819–828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Caamano, J.; Hunter, C.A. NF-κB family of transcription factors: Central regulators of innate and adaptive immune functions. Clin. Microbiol. Rev. 2002, 15, 414–429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Sif, S.; Gilmore, T.D. NF-kappa B p100 is one of the high-molecular-weight proteins complexed with the v-Rel oncoprotein in transformed chicken spleen cells. J. Virol. 1993, 67, 7612–7617. [Google Scholar] [CrossRef] [PubMed]
  8. Ryseck, R.-P.; Bull, P.; Takamiya, M.; Bours, V.; Siebenlist, U.; Dobrzanski, P.; Bravo, R. RelB, a new Rel family transcription activator that can interact with p50-NF-kappa B. Mol. Cell. Biol. 1992, 12, 674–684. [Google Scholar] [PubMed] [Green Version]
  9. Chen, I.; Wilhelmsen, K.; Temin, H.M. Structure and expression of c-rel, the cellular homolog to the oncogene of reticuloendotheliosis virus strain T. J. Virol. 1983, 45, 104–113. [Google Scholar] [CrossRef] [Green Version]
  10. Wilhelmsen, K.C.; Eggleton, K.; Temin, H.M. Nucleic acid sequences of the oncogene v-rel in reticuloendotheliosis virus strain T and its cellular homolog, the proto-oncogene c-rel. J. Virol. 1984, 52, 172–182. [Google Scholar] [CrossRef] [Green Version]
  11. Ghosh, S.; Gifford, A.M.; Riviere, L.R.; Tempst, P.; Nolan, G.P.; Baltimore, D. Cloning of the p50 DNA binding subunit of NF-κB: Homology to rel and dorsal. Cell 1990, 62, 1019–1029. [Google Scholar] [CrossRef] [PubMed]
  12. Ghosh, G.; Duyne, G.V.; Ghosh, S.; Sigler, P.B. Structure of NF-κB p50 homodimer bound to a κB site. Nature 1995, 373, 303–310. [Google Scholar] [CrossRef] [PubMed]
  13. Bours, V.; Villalobos, J.; Burd, P.R.; Kelly, K.; Siebenlist, U. Cloning of a mitogen-inducible gene encoding a κB DNA-binding protein with homology to the rel oncogene and to cell-cycle motifs. Nature 1990, 348, 76–80. [Google Scholar] [CrossRef] [PubMed]
  14. Meyer, R.; Hatada, E.N.; Hohmann, H.-P.; Haiker, M.; Bartsch, C.; Röthlisberger, U.; Lahm, H.-W.; Schlaeger, E.J.; Van Loon, A.; Scheidereit, C. Cloning of the DNA-binding subunit of human nuclear factor kappa B: The level of its mRNA is strongly regulated by phorbol ester or tumor necrosis factor alpha. Proc. Natl. Acad. Sci. USA 1991, 88, 966–970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Neri, A.; Chang, C.-C.; Lombardi, L.; Salina, M.; Corradini, P.; Maiolo, A.T.; Chaganti, R.; Dalla-Favera, R. B cell lymphoma-associated chromosomal translocation involves candidate oncogene lyt-10, homologous to NF-κB p50. Cell 1991, 67, 1075–1087. [Google Scholar] [CrossRef] [PubMed]
  16. Schmid, R.M.; Liptay, S.; Betts, J.C.; Nabel, G.J. Structural and functional analysis of NF-kappa B. Determinants of DNA binding specificity and protein interaction. J. Biol. Chem. 1994, 269, 32162–32167. [Google Scholar] [CrossRef]
  17. MERCURIO, F.; DIDONATO, J.; ROSETTE, C.; Karin, M. Molecular cloning and characterization of a novel rel/NF-χB family member displaying structural and functional homology to NF-χB p50/p105. DNA Cell Biol. 1992, 11, 523–537. [Google Scholar] [CrossRef] [PubMed]
  18. Bours, V.; Burd, P.; Brown, K.; Villalobos, J.; Park, S.; Ryseck, R.P.; Bravo, R.; Kelly, K.; Siebenlist, U. A novel mitogen-inducible gene product related to p50/p105-NF-kappa B participates in transactivation through a kappa B site. Mol. Cell. Biol. 1992, 12, 685–695. [Google Scholar]
  19. Ruben, S.M.; Dillon, P.J.; Schreck, R.; Henkel, T.; Chen, C.-H.; Maher, M.; Baeuerle, P.A.; Rosen, C.A. Isolation of a rel-related Human cDNAThat Potentially Encodes the 65-kD Subunit of NF-κB. Science 1991, 251, 1490–1493. [Google Scholar] [CrossRef]
  20. Ryseck, R.-P.; Novotny, J.; Bravo, R. Characterization of elements determining the dimerization properties of RelB and p50. Mol. Cell. Biol. 1995, 15, 3100–3109. [Google Scholar] [CrossRef] [Green Version]
  21. Bours, V.; Azarenko, V.; Dejardin, E.; Siebenlist, U. Human RelB (I-Rel) functions as a kappa B site-dependent transactivating member of the family of Rel-related proteins. Oncogene 1994, 9, 1699–1702. [Google Scholar]
  22. Inoue, J.-I.; Kerr, L.D.; Kakizuka, A.; Verma, I.M. IκBγ, a 70 kd protein identical to the C-terminal half of p110 NF-κB: A new member of the IκB family. Cell 1992, 68, 1109–1120. [Google Scholar] [CrossRef]
  23. Gerondakis, S.; Morrice, N.; Richardson, I.; Wettenhall, R.; Fecondo, J.; Grumont, R. The activity of a 70 kilodalton I kappa B molecule identical to the carboxyl terminus of the p105 NF-kappa B precursor is modulated by protein kinase A. Cell Growth Differ. Mol. Biol. J. Am. Assoc. Cancer Res. 1993, 4, 617–627. [Google Scholar]
  24. Grumont, R.J.; Gerondakis, S. Alternative splicing of RNA transcripts encoded by the murine p105 NF-kappa B gene generates I kappa B gamma isoforms with different inhibitory activities. Proc. Natl. Acad. Sci. USA 1994, 91, 4367–4371. [Google Scholar] [CrossRef] [Green Version]
  25. Scheidereit, C. IκB kinase complexes: Gateways to NF-κB activation and transcription. Oncogene 2006, 25, 6685–6705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Zandi, E.; Rothwarf, D.M.; Delhase, M.; Hayakawa, M.; Karin, M. The IkB kinase complex (IKK) contains two kinase subunits, IKKa and IKKb, necessary for IkB phosphorylation and NF-kB activation. Cell 1997, 91, 243–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Rothwarf, D.M.; Zandi, E.; Natoli, G.; Karin, M. IKK-γ is an essential regulatory subunit of the IκB kinase complex. Nature 1998, 395, 297–300. [Google Scholar] [CrossRef] [PubMed]
  28. Gilmore, T.D. The Rel/NF-κB signal transduction pathway: Introduction. Oncogene 1999, 18, 6842–6844. [Google Scholar] [CrossRef] [Green Version]
  29. Shindo, M.; Nakano, H.; Sakon, S.; Yagita, H. Assignment of (I kappa B) kinase (beta)(IKBKB) to human chromosome band 8p12--> p11 by in situ hybridization. Cytogenet. Genome Res. 1998, 82, 32. [Google Scholar] [CrossRef]
  30. Ling, L.; Cao, Z.; Goeddel, D.V. NF-κB-inducing kinase activates IKK-α by phosphorylation of Ser-176. Proc. Natl. Acad. Sci. USA 1998, 95, 3792–3797. [Google Scholar] [CrossRef] [Green Version]
  31. Zabel, U.; Henkel, T.; Silva, M.S.; Baeuerle, P.A. Nuclear uptake control of NF-kappa B by MAD-3, an I kappa B protein present in the nucleus. EMBO J. 1993, 12, 201–211. [Google Scholar] [CrossRef]
  32. Zabel, U.; Baeuerle, P.A. Purified human IκB can rapidly dissociate the complex of the NF-κB transcription factor with its cognate DNA. Cell 1990, 61, 255–265. [Google Scholar] [CrossRef]
  33. Arenzana-Seisdedos, F.; Turpin, P.; Rodriguez, M.; Thomas, D.; Hay, R.T.; Virelizier, J.-L.; Dargemont, C. Nuclear localization of I kappa B alpha promotes active transport of NF-kappa B from the nucleus to the cytoplasm. J. Cell Sci. 1997, 110, 369–378. [Google Scholar] [CrossRef]
  34. Zhang, Q.; Didonato, J.A.; Karin, M.; McKeithan, T.W. BCL3 encodes a nuclear protein which can alter the subcellular location of NF-kappa B proteins. Mol. Cell. Biol. 1994, 14, 3915–3926. [Google Scholar] [PubMed] [Green Version]
  35. Hatada, E.N.; Nieters, A.; Wulczyn, F.G.; Naumann, M.; Meyer, R.; Nucifora, G.; McKeithan, T.W.; Scheidereit, C. The ankyrin repeat domains of the NF-kappa B precursor p105 and the protooncogene bcl-3 act as specific inhibitors of NF-kappa B DNA binding. Proc. Natl. Acad. Sci. USA 1992, 89, 2489–2493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Carmody, R.J.; Ruan, Q.; Palmer, S.; Hilliard, B.; Chen, Y.H. Negative regulation of toll-like receptor signaling by NF-κB p50 ubiquitination blockade. Science 2007, 317, 675–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Wang, V.Y.-F.; Huang, W.; Asagiri, M.; Spann, N.; Hoffmann, A.; Glass, C.; Ghosh, G. The transcriptional specificity of NF-κB dimers is coded within the κB DNA response elements. Cell Rep. 2012, 2, 824–839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Pan, W.; Deng, L.; Wang, H.; Wang, V.Y.-F. Atypical IκB Bcl3 enhances the generation of the NF-κB p52 homodimer. Front. Cell Dev. Biol. 2022, 10, 930619. [Google Scholar] [CrossRef] [PubMed]
  39. Totzke, G.; Essmann, F.; Pohlmann, S.; Lindenblatt, C.; Jänicke, R.U.; Schulze-Osthoff, K. A novel member of the IκB family, human IκB-ζ, inhibits transactivation of p65 and its DNA binding. J. Biol. Chem. 2006, 281, 12645–12654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Yamauchi, S.; Ito, H.; Miyajima, A. IκBη, a nuclear IκB protein, positively regulates the NF-κB–mediated expression of proinflammatory cytokines. Proc. Natl. Acad. Sci. USA 2010, 107, 11924–11929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Fiorini, E.; Schmitz, I.; Marissen, W.E.; Osborn, S.L.; Touma, M.; Sasada, T.; Reche, P.A.; Tibaldi, E.V.; Hussey, R.E.; Kruisbeek, A.M. Peptide-induced negative selection of thymocytes activates transcription of an NF-κB inhibitor. Mol. Cell 2002, 9, 637–648. [Google Scholar] [CrossRef]
  42. Schuster, M.; Glauben, R.; Plaza-Sirvent, C.; Schreiber, L.; Annemann, M.; Floess, S.; Kühl, A.A.; Clayton, L.K.; Sparwasser, T.; Schulze-Osthoff, K. IκBNS protein mediates regulatory T cell development via induction of the Foxp3 transcription factor. Immunity 2012, 37, 998–1008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Bubici, C.; Papa, S.; Dean, K.; Franzoso, G. Mutual cross-talk between reactive oxygen species and nuclear factor-kappa B: Molecular basis and biological significance. Oncogene 2006, 25, 6731–6748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Barré, B.; Coqueret, O.; Perkins, N.D. Regulation of activity and function of the p52 NF-κB subunit following DNA damage. Cell Cycle 2010, 9, 4795–4804. [Google Scholar] [CrossRef] [PubMed]
  45. Morgan, M.J.; Liu, Z.-g. Crosstalk of reactive oxygen species and NF-κB signaling. Cell Res. 2011, 21, 103–115. [Google Scholar] [CrossRef] [Green Version]
  46. Osborn, L.; Kunkel, S.; Nabel, G.J. Tumor necrosis factor alpha and interleukin 1 stimulate the human immunodeficiency virus enhancer by activation of the nuclear factor kappa B. Proc. Natl. Acad. Sci. USA 1989, 86, 2336–2340. [Google Scholar] [CrossRef] [Green Version]
  47. Duh, E.J.; Maury, W.J.; Folks, T.M.; Fauci, A.S.; Rabson, A.B. Tumor necrosis factor alpha activates human immunodeficiency virus type 1 through induction of nuclear factor binding to the NF-kappa B sites in the long terminal repeat. Proc. Natl. Acad. Sci. USA 1989, 86, 5974–5978. [Google Scholar] [CrossRef] [Green Version]
  48. Lowenthal, J.W.; Ballard, D.W.; Böhnlein, E.; Greene, W.C. Tumor necrosis factor alpha induces proteins that bind specifically to kappa B-like enhancer elements and regulate interleukin 2 receptor alpha-chain gene expression in primary human T lymphocytes. Proc. Natl. Acad. Sci. USA 1989, 86, 2331–2335. [Google Scholar] [CrossRef] [Green Version]
  49. Awane, M.; Andres, P.G.; Li, D.J.; Reinecker, H.-C. NF-κB-inducing kinase is a common mediator of IL-17-, TNF-α-, and IL-1β-induced chemokine promoter activation in intestinal epithelial cells. J. Immunol. 1999, 162, 5337–5344. [Google Scholar] [CrossRef]
  50. Elewaut, D.; DiDonato, J.A.; Kim, J.M.; Truong, F.; Eckmann, L.; Kagnoff, M.F. NF-κB is a central regulator of the intestinal epithelial cell innate immune response induced by infection with enteroinvasive bacteria. J. Immunol. 1999, 163, 1457–1466. [Google Scholar] [CrossRef]
  51. Hayden, M.S.; Ghosh, S. Signaling to NF-κB. Genes Dev. Biol. 2004, 18, 2195–2224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Claudio, E.; Brown, K.; Park, S.; Wang, H.; Siebenlist, U. BAFF-induced NEMO-independent processing of NF-κB2 in maturing B cells. Nat. Immunol. 2002, 3, 958–965. [Google Scholar] [CrossRef] [PubMed]
  53. Coope, H.; Atkinson, P.; Huhse, B.; Belich, M.; Janzen, J.; Holman, M.; Klaus, G.; Johnston, L.; Ley, S. CD40 regulates the processing of NF-κB2 p100 to p52. EMBO J. 2002, 21, 5375–5385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Dejardin, E.; Droin, N.M.; Delhase, M.; Haas, E.; Cao, Y.; Makris, C.; Li, Z.-W.; Karin, M.; Ware, C.F.; Green, D.R. The lymphotoxin-β receptor induces different patterns of gene expression via two NF-κB pathways. Immunity 2002, 17, 525–535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Kayagaki, N.; Yan, M.; Seshasayee, D.; Wang, H.; Lee, W.; French, D.M.; Grewal, I.S.; Cochran, A.G.; Gordon, N.C.; Yin, J. BAFF/BLyS receptor 3 binds the B cell survival factor BAFF ligand through a discrete surface loop and promotes processing of NF-κB2. Immunity 2002, 17, 515–524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Senftleben, U.; Cao, Y.; Xiao, G.; Greten, F.R.; Krähn, G.; Bonizzi, G.; Chen, Y.; Hu, Y.; Fong, A.; Sun, S.-C.J.S. Activation by IKKα of a second, evolutionary conserved, NF-κB signaling pathway. Science 2001, 293, 1495–1499. [Google Scholar] [CrossRef] [PubMed]
  57. Xiao, G.; Harhaj, E.W.; Sun, S.-C. NF-κB-inducing kinase regulates the processing of NF-κB2 p100. Mol. Cell 2001, 7, 401–409. [Google Scholar] [CrossRef]
  58. Xiao, G.; Fong, A.; Sun, S.-C. Induction of p100 processing by NF-κB-inducing kinase involves docking IκB kinase α (IKKα) to p100 and IKKα-mediated phosphorylation. J. Biol. Chem. 2004, 279, 30099–30105. [Google Scholar] [CrossRef] [Green Version]
  59. Xiao, G.; Cvijic, M.E.; Fong, A.; Harhaj, E.W.; Uhlik, M.T.; Waterfield, M.; Sun, S.C. Retroviral oncoprotein Tax induces processing of NF-κB2/p100 in T cells: Evidence for the involvement of IKKα. EMBO J. 2001, 20, 6805–6815. [Google Scholar] [CrossRef] [Green Version]
  60. Ghosh, G.; Wang, V.Y.-F. Origin of the functional distinctiveness of NF-κB/p52. Front. Cell Dev. Biol. 2021, 9, 3338. [Google Scholar] [CrossRef]
  61. Miyamoto, S.; Schmitt, M.J.; Verma, I.M. Qualitative changes in the subunit composition of kappa B-binding complexes during murine B-cell differentiation. Proc. Natl. Acad. Sci. USA 1994, 91, 5056–5060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Lernbecher, T.; Müller, U.; Wirth, T. Distinct NF-κB/Rel transcription factors are responsible for tissue-specific and inducible gene activation. Nature 1993, 365, 767–770. [Google Scholar] [CrossRef] [PubMed]
  63. Müller, C.W.; Rey, F.A.; Sodeoka, M.; Verdine, G.L.; Harrison, S.C. Structure of the NF-κB p50 homodimer bound to DNA. Nature 1995, 373, 311–317. [Google Scholar] [CrossRef]
  64. Cramer, P.; Larson, C.J.; Verdine, G.L.; Müller, C.W. Structure of the human NF-κB p52 homodimer-DNA complex at 2.1 Å resolution. EMBO J. 1997, 16, 7078–7090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Chen, Y.-Q.; Ghosh, S.; Ghosh, G. A novel DNA recognition mode by the NF-κB p65 homodimer. Nat. Struct. Biol. 1998, 5, 67–73. [Google Scholar] [CrossRef]
  66. Huang, D.-B.; Vu, D.; Ghosh, G. NF-κB RelB forms an intertwined homodimer. Structure 2005, 13, 1365–1373. [Google Scholar] [CrossRef] [Green Version]
  67. Driessler, F.; Venstrom, K.; Sabat, R.; Asadullah, K.; Schottelius, A.J.C. Molecular mechanisms of interleukin-10-mediated inhibition of NF-κ B activity: A role for p50. Clin. Exp. Immunol. 2004, 135, 64–73. [Google Scholar] [CrossRef]
  68. Elsharkawy, A.M.; Oakley, F.; Lin, F.; Packham, G.; Mann, D.A.; Mann, J. The NF-κB p50: p50: HDAC-1 repressor complex orchestrates transcriptional inhibition of multiple pro-inflammatory genes. J. Hepatol. 2010, 53, 519–527. [Google Scholar] [CrossRef] [Green Version]
  69. Ernst, M.K.; Dunn, L.L.; Rice, N.R. The PEST-like sequence of I kappa B alpha is responsible for inhibition of DNA binding but not for cytoplasmic retention of c-Rel or RelA homodimers. Mol. Cell. Biol. 1995, 15, 872–882. [Google Scholar] [CrossRef] [Green Version]
  70. Latimer, M.; Ernst, M.K.; Dunn, L.L.; Drutskaya, M.; Rice, N.R. The N-terminal domain of IκBα masks the nuclear localization signal (s) of p50 and c-Rel homodimers. Mol. Cell. Biol. 1998, 18, 2640–2649. [Google Scholar] [CrossRef] [Green Version]
  71. Basak, S.; Shih, V.F.-S.; Hoffmann, A. Generation and activation of multiple dimeric transcription factors within the NF-κB signaling system. Mol. Cell. Biol. 2008, 28, 3139–3150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Chen, F.E.; Huang, D.-B.; Chen, Y.-Q.; Ghosh, G. Crystal structure of p50/p65 heterodimer of transcription factor NF-κB bound to DNA. Nature 1998, 391, 410–413. [Google Scholar] [CrossRef] [PubMed]
  73. Liao, G.; Zhang, M.; Harhaj, E.W.; Sun, S.-C. Regulation of the NF-κB-inducing kinase by tumor necrosis factor receptor-associated factor 3-induced degradation. J. Biol. Chem. 2004, 279, 26243–26250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. de Jong, S.J.; Albrecht, J.-C.; Giehler, F.; Kieser, A.; Sticht, H.; Biesinger, B. Noncanonical NF-κB activation by the oncoprotein Tio occurs through a nonconserved TRAF3-binding motif. Sci. Signal. 2013, 6, ra27. [Google Scholar] [CrossRef]
  75. He, J.Q.; Zarnegar, B.; Oganesyan, G.; Saha, S.K.; Yamazaki, S.; Doyle, S.E.; Dempsey, P.W.; Cheng, G. Rescue of TRAF3-null mice by p100 NF-κB deficiency. J. Exp. Med. 2006, 203, 2413–2418. [Google Scholar] [CrossRef] [Green Version]
  76. Zarnegar, B.J.; Wang, Y.; Mahoney, D.J.; Dempsey, P.W.; Cheung, H.H.; He, J.; Shiba, T.; Yang, X.; Yeh, W.-c.; Mak, T.W. Noncanonical NF-κB activation requires coordinated assembly of a regulatory complex of the adaptors cIAP1, cIAP2, TRAF2 and TRAF3 and the kinase NIK. Nat. Immunol. 2008, 9, 1371–1378. [Google Scholar] [CrossRef] [Green Version]
  77. Vince, J.E.; Wong, W.W.-L.; Khan, N.; Feltham, R.; Chau, D.; Ahmed, A.U.; Benetatos, C.A.; Chunduru, S.K.; Condon, S.M.; McKinlay, M. IAP antagonists target cIAP1 to induce TNFα-dependent apoptosis. Cell 2007, 131, 682–693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Varfolomeev, E.; Blankenship, J.W.; Wayson, S.M.; Fedorova, A.V.; Kayagaki, N.; Garg, P.; Zobel, K.; Dynek, J.N.; Elliott, L.O.; Wallweber, H.J. IAP antagonists induce autoubiquitination of c-IAPs, NF-κB activation, and TNFα-dependent apoptosis. Cell 2007, 131, 669–681. [Google Scholar] [CrossRef] [Green Version]
  79. Brown, K.D.; Hostager, B.S.; Bishop, G.A. Differential signaling and tumor necrosis factor receptor–associated factor (Traf) degradation mediated by Cd40 and the Epstein-Barr virus oncoprotein latent membrane protein 1 (Lmp1). J. Exp. Med. 2001, 193, 943–954. [Google Scholar] [CrossRef] [Green Version]
  80. Vallabhapurapu, S.; Matsuzawa, A.; Zhang, W.; Tseng, P.-H.; Keats, J.J.; Wang, H.; Vignali, D.A.; Bergsagel, P.L.; Karin, M. Nonredundant and complementary functions of TRAF2 and TRAF3 in a ubiquitination cascade that activates NIK-dependent alternative NF-κB signaling. Nat. Immunol. 2008, 9, 1364–1370. [Google Scholar] [CrossRef] [Green Version]
  81. Liang, C.; Zhang, M.; Sun, S.-C. β-TrCP binding and processing of NF-κB2/p100 involve its phosphorylation at serines 866 and 870. Cell. Signal. 2006, 18, 1309–1317. [Google Scholar] [CrossRef] [PubMed]
  82. Savinova, O.V.; Hoffmann, A.; Ghosh, G. The Nfkb1 and Nfkb2 proteins p105 and p100 function as the core of high-molecular-weight heterogeneous complexes. Mol. Cell 2009, 34, 591–602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Betts, J.C.; Nabel, G.J. Differential regulation of NF-kappaB2 (p100) processing and control by amino-terminal sequences. Mol. Cell. Biol. 1996, 16, 6363–6371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Biswas, S.K.; Lewis, C.E. NF-κB as a central regulator of macrophage function in tumors. J. Leukoc. Biol. 2010, 88, 877–884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Dejardin, E. The alternative NF-κB pathway from biochemistry to biology: Pitfalls and promises for future drug development. J Biochem. Pharmacol. 2006, 72, 1161–1179. [Google Scholar] [CrossRef]
  86. Espinosa, L.; Bigas, A.; Mulero, M.C. Alternative nuclear functions for NF-κB family members. Am. J. Cancer Res. 2011, 1, 446. [Google Scholar] [PubMed]
  87. Hayden, M.S.; Ghosh, S. NF-κB, the first quarter-century: Remarkable progress and outstanding questions. Genes 2012, 26, 203–234. [Google Scholar] [CrossRef] [Green Version]
  88. Hoesel, B.; Schmid, J.A. The complexity of NF-κB signaling in inflammation and cancer. Mol. Cancer Res. 2013, 12, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Oeckinghaus, A.; Ghosh, S. The NF-κB family of transcription factors and its regulation. Cold Spring Harb. Perspect. Biol. 2009, 1, a000034. [Google Scholar] [CrossRef] [PubMed]
  90. Sun, S.-C. Non-canonical NF-κB signaling pathway. Cell Res. 2011, 21, 71–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Keats, J.J.; Fonseca, R.; Chesi, M.; Schop, R.; Baker, A.; Chng, W.-J.; Van Wier, S.; Tiedemann, R.; Shi, C.-X.; Sebag, M. Promiscuous mutations activate the noncanonical NF-κB pathway in multiple myeloma. Cancer Cell 2007, 12, 131–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Yılmaz, Z.B.; Kofahl, B.; Beaudette, P.; Baum, K.; Ipenberg, I.; Weih, F.; Wolf, J.; Dittmar, G.; Scheidereit, C. Quantitative dissection and modeling of the NF-κB p100-p105 module reveals interdependent precursor proteolysis. Cell Rep. 2014, 9, 1756–1769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Annunziata, C.M.; Davis, R.E.; Demchenko, Y.; Bellamy, W.; Gabrea, A.; Zhan, F.; Lenz, G.; Hanamura, I.; Wright, G.; Xiao, W. Frequent engagement of the classical and alternative NF-κB pathways by diverse genetic abnormalities in multiple myeloma. Cancer Cell 2007, 12, 115–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Chapman, M.A.; Lawrence, M.S.; Keats, J.J.; Cibulskis, K.; Sougnez, C.; Schinzel, A.C.; Harview, C.L.; Brunet, J.-P.; Ahmann, G.J.; Adli, M. Initial genome sequencing and analysis of multiple myeloma. Nature 2011, 471, 467–472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Zhang, B.; Calado, D.P.; Wang, Z.; Fröhler, S.; Köchert, K.; Qian, Y.; Koralov, S.B.; Schmidt-Supprian, M.; Sasaki, Y.; Unitt, C. An oncogenic role for alternative NF-κB signaling in DLBCL revealed upon deregulated BCL6 expression. Cell Rep. 2015, 11, 715–726. [Google Scholar] [CrossRef] [Green Version]
  96. SHELDON, W.H.; JAMES, D.F. Cirrhosis following infectious hepatitis: A report of five cases, in two of which there was superimposed primary liver cell carcinoma. Arch. Intern. Med. 1948, 81, 666–689. [Google Scholar] [CrossRef]
  97. Walshe, J.; Wolff, H. Primary Carcinoma of the Liver following Viral Hepatitis. Report of Two Cases. Lancet 1952, 2, 1007–1010. [Google Scholar] [CrossRef]
  98. Steiner, P.; Davies, J. Cirrhosis and primary liver carcinoma in Uganda Africans. Br. J. Cancer 1957, 11, 523. [Google Scholar] [CrossRef] [Green Version]
  99. Higginson, J. The geographical pathology of primary liver cancer. Cancer Res. 1963, 23, 1624–1633. [Google Scholar]
  100. Pisani, P.; Parkin, D.M.; Muñoz, N.; Ferlay, J.J.C.E. Cancer and infection: Estimates of the attributable fraction in 1990. Cancer Epidemiol. Prev. Biomark. 1997, 6, 387–400. [Google Scholar]
  101. Kuper, H.; Adami, H.O.; Trichopoulos, D. Infections as a major preventable cause of human cancer. J. Intern. Med. 2001, 249, 61–74. [Google Scholar] [CrossRef]
  102. Gilmore, T.D.; Gerondakis, S. The c-Rel transcription factor in development and disease. Genes Cancer Cell 2011, 2, 695–711. [Google Scholar] [CrossRef] [Green Version]
  103. Gilmore, T.D. The Rel/NF-κB/IκB signal transduction pathway and cancer. In Signal Transduction Cancer; Spinger: Berlin/Heidelberg, Germany, 2004; pp. 241–265. [Google Scholar] [CrossRef]
  104. Shattuck-Brandt, R.L.; Richmond, A. Enhanced degradation of I-κBα contributes to endogenous activation of NF-κB in Hs294T melanoma cells. Cancer Res. 1997, 57, 3032–3039. [Google Scholar] [PubMed]
  105. Charalambous, M.; Lightfoot, T.; Speirs, V.; Horgan, K.; Gooderham, N. Expression of COX-2, NF-κB-p65, NF-κB-p50 and IKKα in malignant and adjacent normal human colorectal tissue. Br. J. Cancer 2009, 101, 106–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Cogswell, P.C.; Guttridge, D.C.; Funkhouser, W.K.; Baldwin, A.S. Selective activation of NF-κB subunits in human breast cancer: Potential roles for NF-κB2/p52 and for Bcl-3. Oncogene 2000, 19, 1123–1131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Kendellen, M.F.; Bradford, J.W.; Lawrence, C.L.; Clark, K.S.; Baldwin, A.S. Canonical and non-canonical NF-κB signaling promotes breast cancer tumor-initiating cells. Oncogene 2014, 33, 1297–1305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Poma, P.; Labbozzetta, M.; D’Alessandro, N.; Notarbartolo, M. NF-κB is a potential molecular drug target in triple-negative breast cancers. Omics: A J. Integr. Biol. 2017, 21, 225–231. [Google Scholar] [CrossRef]
  109. Devanaboyina, M.; Kaur, J.; Whiteley, E.; Lin, L.; Einloth, K.; Morand, S.; Stanbery, L.; Hamouda, D.; Nemunaitis, J. NF-κB Signaling in Tumor Pathways Focusing on Breast and Ovarian Cancer. Oncol. Rev. 2022, 16, 10568. [Google Scholar] [CrossRef]
  110. Saitoh, Y.; Bruyn, V.J.M.; Uota, S.; Hasegawa, A.; Yamamoto, N.; Imoto, I.; Inazawa, J.; Yamaoka, S. Overexpression of NF-κB inducing kinase underlies constitutive NF-κB activation in lung cancer cells. Lung Cancer 2010, 70, 263–270. [Google Scholar] [CrossRef]
  111. Yang, J.; Splittgerber, R.; Yull, F.E.; Kantrow, S.; Ayers, G.D.; Karin, M.; Richmond, A. Conditional ablation of Ikkb inhibits melanoma tumor development in mice. J. Clin. Investig. 2010, 120, 2563–2574. [Google Scholar] [CrossRef]
  112. Davis, R.E.; Brown, K.D.; Siebenlist, U.; Staudt, L.M. Constitutive nuclear factor κB activity is required for survival of activated B cell–like diffuse large B cell lymphoma cells. J. Exp. Med. 2001, 194, 1861–1874. [Google Scholar] [CrossRef] [PubMed]
  113. Eluard, B.; Nuan-Aliman, S.; Faumont, N.; Collares, D.; Bordereaux, D.; Montagne, A.; Martins, I.; Cagnard, N.; Caly, M.; Taoui, O. The alternative RelB NF-κB subunit is a novel critical player in diffuse large B-cell lymphoma. Blood J. Am. Soc. Hematol. 2022, 139, 384–398. [Google Scholar] [CrossRef] [PubMed]
  114. Kojima, M.; Morisaki, T.; Sasaki, N.; Nakano, K.; Mibu, R.; Tanaka, M.; Katano, M. Increased nuclear factor-kB activation in human colorectal carcinoma and its correlation with tumor progression. Anticancer Res. 2004, 24, 675–682. [Google Scholar] [PubMed]
  115. Li, W.; Tan, D.; Zenali, M.J.; Brown, R.E. Constitutive activation of nuclear factor-kappa B (NF-kB) signaling pathway in fibrolamellar hepatocellular carcinoma. Int. J. Clin. Exp. Pathol. 2010, 3, 238. [Google Scholar]
  116. Lee, D.W.; Ramakrishnan, D.; Valenta, J.; Parney, I.F.; Bayless, K.J.; Sitcheran, R. The NF-κB RelB protein is an oncogenic driver of mesenchymal glioma. PLoS ONE 2013, 8, e57489. [Google Scholar] [CrossRef] [PubMed]
  117. Zeng, F.; Wang, K.; Huang, R.; Liu, Y.; Zhang, Y.; Hu, H. RELB: A novel prognostic marker for glioblastoma as identified by population-based analysis. Oncol. Lett. 2019, 18, 386–394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Cormier, F.; Monjanel, H.; Fabre, C.; Billot, K.; Sapharikas, E.; Chereau, F.; Bordereaux, D.; Molina, T.J.; Avet-Loiseau, H.; Baud, V. Frequent engagement of RelB activation is critical for cell survival in multiple myeloma. PLoS ONE 2013, 8, e59127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Demchenko, Y.N.; Glebov, O.K.; Zingone, A.; Keats, J.J.; Bergsagel, P.L.; Kuehl, W.M. Classical and/or alternative NF-κB pathway activation in multiple myeloma. Blood J. Am. Soc. Hematol. 2010, 115, 3541–3552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Joos, S.; Otaño-Joos, M.I.; Ziegler, S.; Bruderlein, S.; Du Manoir, S.; Bentz, M.; Moller, P.; Lichter, P. Primary mediastinal (thymic) B-cell lymphoma is characterized by gains of chromosomal material including 9p and amplification of the REL gene. Blood 1996, 87, 1571–1578. [Google Scholar] [CrossRef] [Green Version]
  121. Barth, T.F.; Martin-Subero, J.I.; Joos, S.; Menz, C.K.; Hasel, C.; Mechtersheimer, G.; Parwaresch, R.M.; Lichter, P.; Siebert, R.; Möller, P. Gains of 2p involving the REL locus correlate with nuclear c-Rel protein accumulation in neoplastic cells of classical Hodgkin lymphoma. Blood J. Am. Soc. Hematol. 2003, 101, 3681–3686. [Google Scholar] [CrossRef]
  122. Rosenwald, A.; Wright, G.; Chan, W.C.; Connors, J.M.; Campo, E.; Fisher, R.I.; Gascoyne, R.D.; Muller-Hermelink, H.K.; Smeland, E.B.; Giltnane, J.M. The use of molecular profiling to predict survival after chemotherapy for diffuse large-B-cell lymphoma. N. Engl. J. Med. 2002, 346, 1937–1947. [Google Scholar] [CrossRef] [PubMed]
  123. Cherry, E.; Lee, D.; Jung, J.; Sitcheran, R. AI-06NON-CANONICAL NF-kB SIGNALING DRIVES THE AGGRESSIVE INVASIVENESS OF GLIOBLASTOMA. Neuro-Oncol. 2014, 16, v2. [Google Scholar] [CrossRef] [Green Version]
  124. Bours, V.; Dejardin, E.; Goujon-Letawe, F.; Merville, M.-P.; Castronovo, V. The NF-κB transcription factor and cancer: High expression of NF-κB-and IκB-related proteins in tumor cell lines. Biochem. Pharmacol. 1994, 47, 145–149. [Google Scholar] [CrossRef] [PubMed]
  125. Bours, V.; Franzoso, G.; Azarenko, V.; Park, S.; Kanno, T.; Brown, K.; Siebenlist, U. The oncoprotein Bcl-3 directly transactivates through κB motifs via association with DNA-binding p50B homodimers. Cell 1993, 72, 729–739. [Google Scholar] [CrossRef]
  126. Krappmann, D.; Emmerich, F.; Kordes, U.; Scharschmidt, E.; Dörken, B.; Scheidereit, C. Molecular mechanisms of constitutive NF-κB/Rel activation in Hodgkin/Reed-Sternberg cells. Oncogene 1999, 18, 943–953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Keller, S.A.; Schattner, E.J.; Cesarman, E. Inhibition of NF-κB induces apoptosis of KSHV-infected primary effusion lymphoma cells. Blood J. Am. Soc. Hematol. 2000, 96, 2537–2542. [Google Scholar]
  128. Ballard, D.W.; Bohnlein, E.; Lowenthal, J.W.; Wano, Y.; Franza, B.R.; Greene, W.C. HTLV-I tax induces cellular proteins that activate the κB element in the IL-2 receptor α gene. Science 1988, 241, 1652–1655. [Google Scholar] [CrossRef]
  129. Kaltschmidt, C.; Banz-Jansen, C.; Benhidjeb, T.; Beshay, M.; Förster, C.; Greiner, J.; Hamelmann, E.; Jorch, N.; Mertzlufft, F.; Pfitzenmaier, J. A role for NF-κB in organ specific cancer and cancer stem cells. Cancers 2019, 11, 655. [Google Scholar] [CrossRef] [Green Version]
  130. Watnick, R.S. The role of the tumor microenvironment in regulating angiogenesis. Cold Spring Harb. Perspect. Med. 2012, 2, a006676. [Google Scholar] [CrossRef]
  131. Zamarron, B.F.; Chen, W. Dual roles of immune cells and their factors in cancer development and progression. Int. J. Biol. Sci. 2011, 7, 651. [Google Scholar] [CrossRef]
  132. Movahedi, K.; Laoui, D.; Gysemans, C.; Baeten, M.; Stangé, G.; Van den Bossche, J.; Mack, M.; Pipeleers, D.; In’t Veld, P.; De Baetselier, P. Different tumor microenvironments contain functionally distinct subsets of macrophages derived from Ly6C (high) monocytes. Cancer Res. 2010, 70, 5728–5739. [Google Scholar] [CrossRef] [Green Version]
  133. Hagemann, T.; Lawrence, T.; McNeish, I.; Charles, K.A.; Kulbe, H.; Thompson, R.G.; Robinson, S.C.; Balkwill, F.R. “Re-educating” tumor-associated macrophages by targeting NF-κB. J. Exp. Med. 2008, 205, 1261–1268. [Google Scholar] [CrossRef] [Green Version]
  134. Chefetz, I.; Holmberg, J.C.; Alvero, A.B.; Visintin, I.; Mor, G. Inhibition of Aurora-A kinase induces cell cycle arrest in epithelial ovarian cancer stem cells by affecting NFĸB pathway. Cell Cycle 2011, 10, 2206–2214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Sun, S.-C.; Harhaj, E.W. Receptors and adaptors for NF-kB signaling. In NF-Kb/Rel Transcription Factor Fam; Springer: Berlin/Heidelberg, Germany, 2006; pp. 24–26. [Google Scholar]
  136. Novack, D.V.; Yin, L.; Hagen-Stapleton, A.; Schreiber, R.D.; Goeddel, D.V.; Ross, F.P.; Teitelbaum, S.L. The IκB function of NF-κB2 p100 controls stimulated osteoclastogenesis. J. Exp. Med. 2003, 198, 771–781. [Google Scholar] [CrossRef] [PubMed]
  137. Munroe, M.E.; Bishop, G.A. Role of tumor necrosis factor (TNF) receptor-associated factor 2 (TRAF2) in distinct and overlapping CD40 and TNF receptor 2/CD120b-mediated B lymphocyte activation. J. Biol. Chem. 2004, 279, 53222–53231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Rauert, H.; Wicovsky, A.; Müller, N.; Siegmund, D.; Spindler, V.; Waschke, J.; Kneitz, C.; Wajant, H. Membrane tumor necrosis factor (TNF) induces p100 processing via TNF receptor-2 (TNFR2). J. Biol. Chem. 2010, 285, 7394–7404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Saitoh, T.; Nakayama, M.; Nakano, H.; Yagita, H.; Yamamoto, N.; Yamaoka, S. TWEAK induces NF-κB2 p100 processing and long lasting NF-κB activation. J. Biol. Chem. 2003, 278, 36005–36012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Gelbmann, C.; Leeb, S.; Vogl, D.; Maendel, M.; Herfarth, H.; Schölmerich, J.; Falk, W.; Rogler, G. Inducible CD40 expression mediates NFκB activation and cytokine secretion in human colonic fibroblasts. Gut 2003, 52, 1448–1456. [Google Scholar] [CrossRef] [Green Version]
  141. Liu, Z.; Colpaert, S.; D’Haens, G.R.; Kasran, A.; de Boer, M.; Rutgeerts, P.; Geboes, K.; Ceuppens, J.L. Hyperexpression of CD40 ligand (CD154) in inflammatory bowel disease and its contribution to pathogenic cytokine production. J. Immunol. 1999, 163, 4049–4057. [Google Scholar] [CrossRef]
  142. Danese, S.; Scaldaferri, F.; Vetrano, S.; Stefanelli, T.; Graziani, C.; Repici, A.; Ricci, R.; Straface, G.; Sgambato, A.; Malesci, A. Critical role of the CD40–CD40-ligand pathway in regulating mucosal inflammation-driven angiogenesis in inflammatory bowel disease. Gut 2007, 56, 1248–1256. [Google Scholar] [CrossRef] [Green Version]
  143. Blake, M.L.; Tometsko, M.; Miller, R.; Jones, J.C.; Dougall, W.C. RANK expression on breast cancer cells promotes skeletal metastasis. Clin. Exp. Metastasis 2014, 31, 233–245. [Google Scholar] [CrossRef] [PubMed]
  144. Armstrong, A.P.; Miller, R.E.; Jones, J.C.; Zhang, J.; Keller, E.T.; Dougall, W.C. RANKL acts directly on RANK-expressing prostate tumor cells and mediates migration and expression of tumor metastasis genes. Prostate 2008, 68, 92–104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Rossi, D.; Deaglio, S.; Dominguez-Sola, D.; Rasi, S.; Vaisitti, T.; Agostinelli, C.; Spina, V.; Bruscaggin, A.; Monti, S.; Cerri, M. Alteration of BIRC3 and multiple other NF-κB pathway genes in splenic marginal zone lymphoma. Blood J. Am. Soc. Hematol. 2011, 118, 4930–4934. [Google Scholar]
  146. Hyeon, J.; Lee, B.; Shin, S.-H.; Yoo, H.Y.; Kim, S.J.; Kim, W.S.; Park, W.-Y.; Ko, Y.-H. Targeted deep sequencing of gastric marginal zone lymphoma identified alterations of TRAF3 and TNFAIP3 that were mutually exclusive for MALT1 rearrangement. Mod. Pathol. 2018, 31, 1418–1428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Courtois, G.; Gilmore, T. Mutations in the NF-κ B signaling pathway: Implications for human disease. Oncogene 2006, 25, 6831–6843. [Google Scholar] [CrossRef] [Green Version]
  148. Fracchiolla, N.; Lombardi, L.; Salina, M.; Migliazza, A.; Baldini, L.; Berti, E.; Cro, L.; Polli, E.; Maiolo, A.; Neri, A. Structural alterations of the NF-kappa B transcription factor lyt-10 in lymphoid malignancies. Oncogene 1993, 8, 2839–2845. [Google Scholar]
  149. Barth, T.F.; Döhner, H.; Werner, C.A.; Stilgenbauer, S.; Schlotter, M.; Pawlita, M.; Lichter, P.; Möller, P.; Bentz, M. Characteristic pattern of chromosomal gains and losses in primary large B-cell lymphomas of the gastrointestinal tract. Blood J. Am. Soc. Hematol. 1998, 91, 4321–4330. [Google Scholar]
  150. Lu, D.; Thompson, J.; Gorski, G.; Rice, N.; Mayer, M.; Yunis, J. Alterations at the rel locus in human lymphoma. Oncogene 1991, 6, 1235–1241. [Google Scholar]
  151. Kalaitzidis, D.; Gilmore, T.D. Genomic organization and expression of the rearranged REL proto-oncogene in the human B-cell lymphoma cell line RC-K8. Genes Chromosom. Cancer 2002, 34, 129–135. [Google Scholar] [CrossRef]
  152. Zhang, M.; Xu-Monette, Z.Y.; Li, L.; Manyam, G.C.; Visco, C.; Tzankov, A.; Wang, J.; Montes-Moreno, S.; Dybkaer, K.; Chiu, A. RelA NF-κB subunit activation as a therapeutic target in diffuse large B-cell lymphoma. Aging 2016, 8, 3321. [Google Scholar] [CrossRef] [Green Version]
  153. Willis, T.G.; Jadayel, D.M.; Du, M.-Q.; Peng, H.; Perry, A.R.; Abdul-Rauf, M.; Price, H.; Karran, L.; Majekodunmi, O.; Wlodarska, I.J.C. Bcl10 is involved in t (1; 14)(p22; q32) of MALT B cell lymphoma and mutated in multiple tumor types. Cell 1999, 96, 35–45. [Google Scholar] [CrossRef] [Green Version]
  154. Akagi, T.; Motegi, M.; Tamura, A.; Suzuki, R.; Hosokawa, Y.; Suzuki, H.; Ota, H.; Nakamura, S.; Morishima, Y.; Taniwaki, M. A novel gene, MALT1 at 18q21, is involved in t (11; 18)(q21; q21) found in low-grade B-cell lymphoma of mucosa-associated lymphoid tissue. Oncogene 1999, 18, 5785–5794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Lucas, P.C.; Yonezumi, M.; Inohara, N.; McAllister-Lucas, L.M.; Abazeed, M.E.; Chen, F.F.; Yamaoka, S.; Seto, M.; Núñez, G. Bcl10 and MALT1, independent targets of chromosomal translocation in malt lymphoma, cooperate in a novel NF-κB signaling pathway. J. Biol. Chem. 2001, 276, 19012–19019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Uren, A.G.; O’Rourke, K.; Aravind, L.A.; Pisabarro, M.T.; Seshagiri, S.; Koonin, E.V.; Dixit, V.M. Identification of paracaspases and metacaspases: Two ancient families of caspase-like proteins, one of which plays a key role in MALT lymphoma. Mol. Cell 2000, 6, 961–967. [Google Scholar] [CrossRef] [PubMed]
  157. Chang, C.-C.; Zhang, J.; Lombardi, L.; Neri, A.; Dalla-Favera, R. Rearranged NFKB-2 genes in lymphoid neoplasms code for constitutively active nuclear transactivators. Mol. Cell. Biol. 1995, 15, 5180–5187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Ishikawa, H.; Carrasco, D.; Claudio, E.; Ryseck, R.-P.; Bravo, R. Gastric hyperplasia and increased proliferative responses of lymphocytes in mice lacking the COOH-terminal ankyrin domain of NF-κB2. J. Exp. Med. 1997, 186, 999–1014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Houldsworth, J.; Mathew, S.; Rao, P.H.; Dyomina, K.; Louie, D.C.; Parsa, N.; Offit, K.; Chaganti, R. REL proto-oncogene is frequently amplified in extranodal diffuse large cell lymphoma. Blood 1996, 87, 25–29. [Google Scholar] [CrossRef] [Green Version]
  160. Kralova, J.; Schatzle, J.; Bargmann, W.; Bose, H. Transformation of avian fibroblasts overexpressing the c-rel proto-oncogene and a variant of c-rel lacking 40 C-terminal amino acids. J. Virol. 1994, 68, 2073–2083. [Google Scholar] [CrossRef] [Green Version]
  161. Rao, P.H.; Houldsworth, J.; Dyomina, K.; Parsa, N.Z.; Cigudosa, J.C.; Louie, D.C.; Popplewell, L.; Offit, K.; Jhanwar, S.C.; Chaganti, R. Chromosomal and gene amplification in diffuse large B-cell lymphoma. J. Am. Soc. Hematol. 1998, 92, 234–240. [Google Scholar] [CrossRef]
  162. Mathew, S.; Murty, V.; Dalla-Favera, R.; Chaganti, R. Chromosomal localization of genes encoding the transcription factors, c-rel, NF-kBp50, NF-cBp65, and lyt-10 by fluorescence in situ. Oncogene 1993, 8, 191–193. [Google Scholar] [PubMed]
  163. Ohno, H.; Takimoto, G.; McKeithan, T.W. The candidate proto-oncogene bcl-3 is related to genes implicated in cell lineage determination and cell cycle control. Cell 1990, 60, 991–997. [Google Scholar] [CrossRef]
  164. Rylatt, D.B.; Aitken, A.; Bilham, T.; Condon, G.D.; Embi, N.; Cohen, P. Glycogen synthase from rabbit skeletal muscle: Amino acid sequence at the sites phosphorylated by glycogen synthase kinase-3, and extension of the N-terminal sequence containing the site phosphorylated by phosphorylase kinase. Eur. J. Biochem. 1980, 107, 529–537. [Google Scholar] [CrossRef]
  165. Kaidanovich-Beilin, O.; Woodgett, J.R. GSK-3: Functional insights from cell biology and animal models. Front. Mol. Neurosci. 2011, 4, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Hoeflich, K.P.; Luo, J.; Rubie, E.A.; Tsao, M.-S.; Jin, O.; Woodgett, J.R. Requirement for glycogen synthase kinase-3β in cell survival and NF-κB activation. Nature 2000, 406, 86–90. [Google Scholar] [CrossRef] [PubMed]
  167. Ougolkov, A.V.; Fernandez-Zapico, M.E.; Savoy, D.N.; Urrutia, R.A.; Billadeau, D.D. Glycogen synthase kinase-3β participates in nuclear factor κB–mediated gene transcription and cell survival in pancreatic cancer cells. Cancer Res. 2005, 65, 2076–2081. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Ougolkov, A.V.; Bone, N.D.; Fernandez-Zapico, M.E.; Kay, N.E.; Billadeau, D.D. Inhibition of glycogen synthase kinase-3 activity leads to epigenetic silencing of nuclear factor κB target genes and induction of apoptosis in chronic lymphocytic leukemia B cells. Blood J. Am. Soc. Hematol. 2007, 110, 735–742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Kotliarova, S.; Pastorino, S.; Kovell, L.C.; Kotliarov, Y.; Song, H.; Zhang, W.; Bailey, R.; Maric, D.; Zenklusen, J.C.; Lee, J. Glycogen synthase kinase-3 inhibition induces glioma cell death through c-MYC, nuclear factor-κB, and glucose regulation. Cancer Res. 2008, 68, 6643–6651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Mazor, M.; Kawano, Y.; Zhu, H.; Waxman, J.; Kypta, R.M. Inhibition of glycogen synthase kinase-3 represses androgen receptor activity and prostate cancer cell growth. Oncogene 2004, 23, 7882–7892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Shakoori, A.; Ougolkov, A.; Yu, Z.W.; Zhang, B.; Modarressi, M.H.; Billadeau, D.D.; Mai, M.; Takahashi, Y.; Minamoto, T. Deregulated GSK3β activity in colorectal cancer: Its association with tumor cell survival and proliferation. Biochem. Biophys. Res. Commun. 2005, 334, 1365–1373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Schulze-Osthoff, K.; Ferrari, D.; Riehemann, K.; Wesselborg, S. Regulation of NF-κB activation by MAP kinase cascades. Immunobiology 1997, 198, 35–49. [Google Scholar] [CrossRef]
  173. De Smaele, E.; Zazzeroni, F.; Papa, S.; Nguyen, D.U.; Jin, R.; Jones, J.; Cong, R.; Franzoso, G. Induction of gadd45 β by NF-κ B downregulates pro-apoptotic JNK signalling. Nature 2001, 414, 308–313. [Google Scholar] [CrossRef] [PubMed]
  174. Papa, S.; Zazzeroni, F.; Bubici, C.; Jayawardena, S.; Alvarez, K.; Matsuda, S.; Nguyen, D.U.; Pham, C.G.; Nelsbach, A.H.; Melis, T. Gadd45β mediates the NF-κB suppression of JNK signalling by targeting MKK7/JNKK2. Nat. Cell Biol. 2004, 6, 146–153. [Google Scholar] [CrossRef] [PubMed]
  175. Vermeulen, L.; De Wilde, G.; Van Damme, P.; Vanden Berghe, W.; Haegeman, G. Transcriptional activation of the NF-κB p65 subunit by mitogen-and stress-activated protein kinase-1 (MSK1). EMBO J. 2003, 22, 1313–1324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Berghe, W.V.; Plaisance, S.; Boone, E.; De Bosscher, K.; Schmitz, M.L.; Fiers, W.; Haegeman, G. p38 and extracellular signal-regulated kinase mitogen-activated protein kinase pathways are required for nuclear factor-κB p65 transactivation mediated by tumor necrosis factor. J. Biol. Chem. 1998, 273, 3285–3290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Bergsagel, P.L.; Carpten, J.; Chesi, M.; VanWier, S.; Keats, J.J.; Sebag, M.; Chng, W.-J.; Schop, R.; Baker, A.; Chng, T.-H. Promiscuous Mutations Frequently Activate the Non-Canonical NFkB Pathway in Multiple Myeloma (MM). Blood 2006, 108, 109. [Google Scholar] [CrossRef]
  178. Liu, Y.; Peng, J.; Sun, T.; Li, N.; Zhang, L.; Ren, J.; Yuan, H.; Kan, S.; Pan, Q.; Li, X. Epithelial EZH2 serves as an epigenetic determinant in experimental colitis by inhibiting TNFα-mediated inflammation and apoptosis. Proc. Natl. Acad. Sci. USA 2017, 114, E3796–E3805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Yamamoto, M.; Ito, T.; Shimizu, T.; Ishida, T.; Semba, K.; Watanabe, S.; Yamaguchi, N.; Inoue, J.i. Epigenetic alteration of the NF-κB-inducing kinase (NIK) gene is involved in enhanced NIK expression in basal-like breast cancer. Cancer Sci. 2010, 101, 2391–2397. [Google Scholar] [CrossRef]
  180. Zanotto-Filho, A.; Goncalves, R.M.; Klafke, K.; de Souza, P.O.; Dillenburg, F.C.; Carro, L.; Gelain, D.P.; Moreira, J.C.F. Inflammatory landscape of human brain tumors reveals an NFκB dependent cytokine pathway associated with mesenchymal glioblastoma. Cancer Lett. 2017, 390, 176–187. [Google Scholar] [CrossRef] [PubMed]
  181. Smale, S.T. Hierarchies of NF-κB target-gene regulation. Nat. Immunol. 2011, 12, 689–694. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Storz, P.; Döppler, H.; Ferran, C.; Grey, S.T.; Toker, A. Functional dichotomy of A20 in apoptotic and necrotic cell death. Biochem. J. 2005, 387, 47–55. [Google Scholar] [CrossRef] [Green Version]
  183. Storz, P.; Döppler, H.; Toker, A. Protein kinase Cδ selectively regulates protein kinase D-dependent activation of NF-κB in oxidative stress signaling. Mol. Cell. Biol. 2004, 24, 2614–2626. [Google Scholar] [CrossRef] [Green Version]
  184. Tanaka, H.; Matsumura, I.; Ezoe, S.; Satoh, Y.; Sakamaki, T.; Albanese, C.; Machii, T.; Pestell, R.G.; Kanakura, Y. E2F1 and c-Myc potentiate apoptosis through inhibition of NF-κB activity that facilitates MnSOD-mediated ROS elimination. Mol. Cell 2002, 9, 1017–1029. [Google Scholar] [CrossRef] [PubMed]
  185. Ji, Z.; He, L.; Regev, A.; Struhl, K. Inflammatory regulatory network mediated by the joint action of NF-kB, STAT3, and AP-1 factors is involved in many human cancers. Proc. Natl. Acad. Sci. USA 2019, 116, 9453–9462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Kay, J.; Thadhani, E.; Samson, L.; Engelward, B. Inflammation-induced DNA damage, mutations and cancer. DNA Repair 2019, 83, 102673. [Google Scholar] [CrossRef] [PubMed]
  187. Liou, G.-Y.; Storz, P. Reactive oxygen species in cancer. Free Radic. Res. 2010, 44, 479–496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Killela, P.J.; Reitman, Z.J.; Jiao, Y.; Bettegowda, C.; Agrawal, N.; Diaz, L.A.; Friedman, A.H.; Friedman, H.; Gallia, G.L.; Giovanella, B.C. TERT promoter mutations occur frequently in gliomas and a subset of tumors derived from cells with low rates of self-renewal. Proc. Natl. Acad. Sci. USA 2013, 110, 6021–6026. [Google Scholar] [CrossRef] [Green Version]
  189. Huang, F.W.; Hodis, E.; Xu, M.J.; Kryukov, G.V.; Chin, L.; Garraway, L.A. Highly recurrent TERT promoter mutations in human melanoma. Science 2013, 339, 957–959. [Google Scholar] [CrossRef] [Green Version]
  190. Lorbeer, F.K.; Hockemeyer, D. TERT promoter mutations and telomeres during tumorigenesis. Curr. Opin. Genet. Dev. Biol. 2020, 60, 56–62. [Google Scholar] [CrossRef]
  191. Horn, S.; Figl, A.; Rachakonda, P.S.; Fischer, C.; Sucker, A.; Gast, A.; Kadel, S.; Moll, I.; Nagore, E.; Hemminki, K. TERT promoter mutations in familial and sporadic melanoma. Science 2013, 339, 959–961. [Google Scholar] [CrossRef] [Green Version]
  192. Li, Y.; Zhou, Q.-L.; Sun, W.; Chandrasekharan, P.; Cheng, H.S.; Ying, Z.; Lakshmanan, M.; Raju, A.; Tenen, D.G.; Cheng, S.-Y. Non-canonical NF-κB signalling and ETS1/2 cooperatively drive C250T mutant TERT promoter activation. Nat. Cell Biol. 2015, 17, 1327–1338. [Google Scholar] [CrossRef]
  193. Xu, X.; Li, Y.; Bharath, S.R.; Ozturk, M.B.; Bowler, M.W.; Loo, B.Z.L.; Tergaonkar, V.; Song, H. Structural basis for reactivating the mutant TERT promoter by cooperative binding of p52 and ETS1. Nat. Commun. 2018, 9, 1–10. [Google Scholar] [CrossRef] [PubMed]
  194. Wang, C.-Y.; Mayo, M.W.; Korneluk, R.G.; Goeddel, D.V.; Baldwin, A.S. NF-κB antiapoptosis: Induction of TRAF1 and TRAF2 and c-IAP1 and c-IAP2 to suppress caspase-8 activation. Science 1998, 281, 1680–1683. [Google Scholar] [CrossRef] [PubMed]
  195. Chen, G.G.; Lee, J.F.; Wang, S.H.; Chan, U.P.; Ip, P.C.; Lau, W.Y. Apoptosis induced by activation of peroxisome-proliferator activated receptor-gamma is associated with Bcl-2 and Nf-kB in human colon cancer. Life Sci. 2002, 70, 2631–2646. [Google Scholar] [CrossRef] [PubMed]
  196. Wu, M.X.; Ao, Z.; Prasad, K.; Wu, R.; Schlossman, S.F. IEX-1L, an apoptosis inhibitor involved in NF-κB-mediated cell survival. Science 1998, 281, 998–1001. [Google Scholar] [CrossRef] [PubMed]
  197. Wang, L.; Du, F.; Wang, X. TNF-α induces two distinct caspase-8 activation pathways. Cell 2008, 133, 693–703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Varfolomeev, E.; Goncharov, T.; Fedorova, A.V.; Dynek, J.N.; Zobel, K.; Deshayes, K.; Fairbrother, W.J.; Vucic, D. c-IAP1 and c-IAP2 are critical mediators of tumor necrosis factor α (TNFα)-induced NF-κB activation. J. Biol. Chem. 2008, 283, 24295–24299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Huber, M.A.; Azoitei, N.; Baumann, B.; Grünert, S.; Sommer, A.; Pehamberger, H.; Kraut, N.; Beug, H.; Wirth, T. NF-κB is essential for epithelial-mesenchymal transition and metastasis in a model of breast cancer progression. J. Clin. Investig. 2004, 114, 569–581. [Google Scholar] [CrossRef] [PubMed]
  200. Yang, J.; Mani, S.A.; Donaher, J.L.; Ramaswamy, S.; Itzykson, R.A.; Come, C.; Savagner, P.; Gitelman, I.; Richardson, A.; Weinberg, R.A. Twist, a master regulator of morphogenesis, plays an essential role in tumor metastasis. Cell 2004, 117, 927–939. [Google Scholar] [CrossRef] [Green Version]
  201. Huber, M.A.; Beug, H.; Wirth, T. Epithelial-mesenchymal transition: NF-κB takes center stage. Cell Cycle 2004, 3, 1477–1480. [Google Scholar] [CrossRef] [Green Version]
  202. Duran, C.; Lee, D.; Jung, J.; Ravi, S.; Pogue, C.; Toussaint, L.; Bayless, K.; Sitcheran, R. NIK regulates MT1-MMP activity and promotes glioma cell invasion independently of the canonical NF-κB pathway. Oncogenesis 2016, 5, e231. [Google Scholar] [CrossRef] [Green Version]
  203. Xie, T.-X.; Xia, Z.; Zhang, N.; Gong, W.; Huang, S. Constitutive NF-κB activity regulates the expression of VEGF and IL-8 and tumor angiogenesis of human glioblastoma. Oncol. Rep. 2010, 23, 725–732. [Google Scholar] [PubMed]
  204. Yoshida, A.; Yoshida, S.; Ishibashi, T.; Kuwano, M.; Inomata, H. Suppression of retinal neovascularization by the NF-kappaB inhibitor pyrrolidine dithiocarbamate in mice. Investig. Ophthalmol. Vis. Sci. 1999, 40, 1624–1629. [Google Scholar]
  205. Pollard, J.W. Tumour-educated macrophages promote tumour progression and metastasis. Nat. Rev. Cancer 2004, 4, 71–78. [Google Scholar] [CrossRef] [PubMed]
  206. Sparmann, A.; Bar-Sagi, D. Ras-induced interleukin-8 expression plays a critical role in tumor growth and angiogenesis. Cancer Cell 2004, 6, 447–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Schmidt, D.; Textor, B.; Pein, O.T.; Licht, A.H.; Andrecht, S.; Sator-Schmitt, M.; Fusenig, N.E.; Angel, P.; Schorpp-Kistner, M. Critical role for NF-κB-induced JunB in VEGF regulation and tumor angiogenesis. EMBO J. 2007, 26, 710–719. [Google Scholar] [CrossRef] [Green Version]
  208. Duyao, M.P.; Buckler, A.J.; Sonenshein, G.E. Interaction of an NF-kappa B-like factor with a site upstream of the c-myc promoter. Proc. Natl. Acad. Sci. USA 1990, 87, 4727–4731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Duyao, M.; Kessler, D.; Spicer, D.; Bartholomew, C.; Cleveland, J.; Siekevitz, M.; Sonenshein, G. Transactivation of the c-myc promoter by human T cell leukemia virus type 1 tax is mediated by NF kappa B. J. Biol. Chem. 1992, 267, 16288–16291. [Google Scholar] [CrossRef]
  210. Kisseleva, T.; Song, L.; Vorontchikhina, M.; Feirt, N.; Kitajewski, J.; Schindler, C. NF-κB regulation of endothelial cell function during LPS-induced toxemia and cancer. J. Clin. Investig. 2006, 116, 2955–2963. [Google Scholar] [CrossRef] [Green Version]
  211. Chan, D.A.; Kawahara, T.L.; Sutphin, P.D.; Chang, H.Y.; Chi, J.-T.; Giaccia, A.J. Tumor vasculature is regulated by PHD2-mediated angiogenesis and bone marrow-derived cell recruitment. Cancer Cell 2009, 15, 527–538. [Google Scholar] [CrossRef] [Green Version]
  212. Rojo, F.; González-Pérez, A.; Furriol, J.; Nicolau, M.J.; Ferrer, J.; Burgués, O.; Sabbaghi, M.; González-Navarrete, I.; Cristobal, I.; Serrano, L. Non-canonical NF-κ B pathway activation predicts outcome in borderline oestrogen receptor positive breast carcinoma. Br. J. Cancer 2016, 115, 322–331. [Google Scholar] [CrossRef]
  213. Espinoza-Sánchez, N.A.; Győrffy, B.; Fuentes-Pananá, E.M.; Götte, M. Differential impact of classical and non-canonical NF-κB pathway-related gene expression on the survival of breast cancer patients. J. Cancer 2019, 10, 5191. [Google Scholar] [CrossRef] [PubMed]
  214. Qin, H.; Zhou, J.; Zhou, P.; Xu, J.; Tang, Z.; Ma, H.; Guo, F. Prognostic significance of RelB overexpression in non-small cell lung cancer patients. Thorac. Cancer 2016, 7, 415–421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Lim, S.K.; Peng, C.C.; Low, S.; Vijay, V.; Budiman, A.; Phang, B.H.; Lim, J.Q.; Jeyasekharan, A.D.; Lim, S.T.; Ong, C.K. Sustained activation of non-canonical NF-κB signalling drives glycolytic reprogramming in doxorubicin-resistant DLBCL. Leukemia 2022, 37, 441–452. [Google Scholar] [CrossRef] [PubMed]
  216. Böhlig, L.; Rother, K. One function—Multiple mechanisms: The manifold activities of p53 as a transcriptional repressor. J. Biomed. Biotechnol. 2011, 2011, 464916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Perkins, N.D. Achieving transcriptional specificity with NF-κB. Int. J. Biochem. Cell Biol. 1997, 29, 1433–1448. [Google Scholar] [CrossRef] [PubMed]
  218. Wang, S.; Liu, Z.; Wang, L.; Zhang, X. NF-κB signaling pathway, inflammation and colorectal cancer. Cell. Mol. Immunol. 2009, 6, 327–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Greten, F.R.; Eckmann, L.; Greten, T.F.; Park, J.M.; Li, Z.-W.; Egan, L.J.; Kagnoff, M.F.; Karin, M. IKKβ links inflammation and tumorigenesis in a mouse model of colitis-associated cancer. Cell 2004, 118, 285–296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Bohuslav, J.; Chen, L.-f.; Kwon, H.; Mu, Y.; Greene, W.C. p53 induces NF-κB activation by an IκB kinase-independent mechanism involving phosphorylation of p65 by ribosomal S6 kinase 1. J. Biol. Chem. 2004, 279, 26115–26125. [Google Scholar] [CrossRef] [Green Version]
  221. Webster, G.A.; Perkins, N.D. Transcriptional cross talk between NF-κB and p53. Mol. Cell. Biol. 1999, 19, 3485–3495. [Google Scholar] [CrossRef] [Green Version]
  222. Vaughan, C.A.; Singh, S.; Windle, B.; Sankala, H.M.; Graves, P.R.; Yeudall, W.A.; Deb, S.P.; Deb, S. p53 mutants induce transcription of NF-κB2 in H1299 cells through CBP and STAT binding on the NF-κB2 promoter and gain of function activity. Arch. Biochem. Biophys. 2012, 518, 79–88. [Google Scholar] [CrossRef] [Green Version]
  223. Chen, L.-F.; Greene, W.C. Regulation of distinct biological activities of the NF-κB transcription factor complex by acetylation. J. Mol. Med. 2003, 81, 549–557. [Google Scholar] [CrossRef] [PubMed]
  224. Kim, J.-W.; Jang, S.-M.; Kim, C.-H.; An, J.-H.; Kang, E.-J.; Choi, K.-H. New molecular bridge between RelA/p65 and NF-κB target genes via histone acetyltransferase TIP60 cofactor. J. Biol. Chem. 2012, 287, 7780–7791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Beg, A.A.; Baltimore, D. An essential role for NF-κB in preventing TNF-α-induced cell death. Science 1996, 274, 782–784. [Google Scholar] [CrossRef] [PubMed]
  226. Schneider, G.; Henrich, A.; Greiner, G.; Wolf, V.; Lovas, A.; Wieczorek, M.; Wagner, T.; Reichardt, S.; Von Werder, A.; Schmid, R. Cross talk between stimulated NF-κB and the tumor suppressor p53. Oncogene 2010, 29, 2795–2806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Ichikawa, H.; Shimizu, K.; Hayashi, Y.; Ohki, M. An RNA-binding protein gene, TLS/FUS, is fused to ERG in human myeloid leukemia with t (16; 21) chromosomal translocation. Cancer Res. 1994, 54, 2865–2868. [Google Scholar]
  228. Prasad, D.; Ouchida, M.; Lee, L.; Rao, V.N.; Reddy, E. TLS/FUS fusion domain of TLS/FUS-erg chimeric protein resulting from the t (16; 21) chromosomal translocation in human myeloid leukemia functions as a transcriptional activation domain. Oncogene 1994, 9, 3717–3729. [Google Scholar]
  229. Sorensen, P.H.; Lessnick, S.L.; Lopez-Terrada, D.; Liu, X.F.; Triche, T.J.; Denny, C.T. A second Ewing’s sarcoma translocation, t (21; 22), fuses the EWS gene to another ETS–family transcription factor, ERG. Nat. Genet. 1994, 6, 146–151. [Google Scholar] [CrossRef]
  230. Tomlins, S.A.; Rhodes, D.R.; Perner, S.; Dhanasekaran, S.M.; Mehra, R.; Sun, X.-W.; Varambally, S.; Cao, X.; Tchinda, J.; Kuefer, R. Recurrent fusion of TMPRSS2 and ETS transcription factor genes in prostate cancer. Science 2005, 310, 644–648. [Google Scholar] [CrossRef]
  231. Dryden, N.H.; Sperone, A.; Martin-Almedina, S.; Hannah, R.L.; Birdsey, G.M.; Khan, S.T.; Layhadi, J.A.; Mason, J.C.; Haskard, D.O.; Göttgens, B. The transcription factor Erg controls endothelial cell quiescence by repressing activity of nuclear factor (NF)-κB p65. J. Biol. Chem. 2012, 287, 12331–12342. [Google Scholar] [CrossRef] [Green Version]
  232. Wang, J.; Cai, Y.; Shao, L.-j.; Siddiqui, J.; Palanisamy, N.; Li, R.; Ren, C.; Ayala, G.; Ittmann, M. Activation of NF-κB by TMPRSS2/ERG fusion isoforms through Toll-like receptor-4. Cancer Res. 2011, 71, 1325–1333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Cai, J.; Kandagatla, P.; Singareddy, R.; Kropinski, A.; Sheng, S.; Cher, M.L.; Chinni, S.R. Androgens induce functional CXCR4 through ERG factor expression in TMPRSS2-ERG fusion-positive prostate cancer cells. Transl. Oncol. 2010, 3, 195-IN191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Sim, N.; Li, Y. NF-κB/p52 augments ETS1 binding genome-wide to promote glioma progression. bioRxiv 2023. bioRxiv 2023.2001. [Google Scholar] [CrossRef]
  235. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Bazzoni, F.; Rossato, M.; Fabbri, M.; Gaudiosi, D.; Mirolo, M.; Mori, L.; Tamassia, N.; Mantovani, A.; Cassatella, M.A.; Locati, M. Induction and regulatory function of miR-9 in human monocytes and neutrophils exposed to proinflammatory signals. Proc. Natl. Acad. Sci. USA 2009, 106, 5282–5287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Fesler, A.; Xu, X.; Zheng, X.; Li, X.; Jiang, J.; Russo, J.J.; Ju, J. Identification of miR-215 mediated targets/pathways via translational immunoprecipitation expression analysis (TrIP-chip). Oncotarget 2015, 6, 24463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Niu, J.; Shi, Y.; Tan, G.; Yang, C.H.; Fan, M.; Pfeffer, L.M.; Wu, Z.-H. DNA Damage Induces NF-κB-Dependent microRNA-21 up-Regulation and Promotes Breast Cancer Cell Invasion. J. Biol. Chem. 2012, 287, 21783–21795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Zhang, X.; Liu, S.; Hu, T.; Liu, S.; He, Y.; Sun, S. Up-regulated microRNA-143 transcribed by nuclear factor kappa B enhances hepatocarcinoma metastasis by repressing fibronectin expression. Hepatology 2009, 50, 490–499. [Google Scholar] [CrossRef]
  240. Taganov, K.D.; Boldin, M.P.; Chang, K.-J.; Baltimore, D. NF-κB-dependent induction of microRNA miR-146, an inhibitor targeted to signaling proteins of innate immune responses. Proc. Natl. Acad. Sci. USA 2006, 103, 12481–12486. [Google Scholar] [CrossRef] [Green Version]
  241. Scisciani, C.; Vossio, S.; Guerrieri, F.; Schinzari, V.; De Iaco, R.; de Meo, P.D.O.; Cervello, M.; Montalto, G.; Pollicino, T.; Raimondo, G. Transcriptional regulation of miR-224 upregulated in human HCCs by NFκB inflammatory pathways. J. Hepatol. 2012, 56, 855–861. [Google Scholar] [CrossRef]
  242. Iliopoulos, D.; Hirsch, H.A.; Struhl, K. An epigenetic switch involving NF-κB, Lin28, Let-7 MicroRNA, and IL6 links inflammation to cell transformation. Cell 2009, 139, 693–706. [Google Scholar] [CrossRef] [Green Version]
  243. Zhang, S.; Shan, C.; Kong, G.; Du, Y.; Ye, L.; Zhang, X. MicroRNA-520e suppresses growth of hepatoma cells by targeting the NF-κB-inducing kinase (NIK). Oncogene 2012, 31, 3607–3620. [Google Scholar] [CrossRef] [Green Version]
  244. Liu, B.; Sun, L.; Liu, Q.; Gong, C.; Yao, Y.; Lv, X.; Lin, L.; Yao, H.; Su, F.; Li, D. A cytoplasmic NF-κB interacting long noncoding RNA blocks IκB phosphorylation and suppresses breast cancer metastasis. Cancer Cell 2015, 27, 370–381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Liao, Z.; Zhao, J.; Yang, Y. Downregulation of lncRNA H19 inhibits the migration and invasion of melanoma cells by inactivating the NF-κB and PI3K/Akt signaling pathways. Mol. Med. Rep. 2018, 17, 7313–7318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Zhang, Y.; Yan, J.; Li, C.; Wang, X.; Dong, Y.; Shen, X.; Zhang, X. LncRNA H19 induced by helicobacter pylori infection promotes gastric cancer cell growth via enhancing NF-κB-induced inflammation. J. Inflamm. 2019, 16, 1–8. [Google Scholar] [CrossRef] [PubMed]
  247. Shen, X.; Zhao, W.; Zhang, Y.; Liang, B. Long non-coding RNA-NEAT1 promotes cell migration and invasion via regulating miR-124/NF-κB pathway in cervical cancer. OncoTargets Ther. 2020, 13, 3265. [Google Scholar] [CrossRef] [Green Version]
  248. Pan, Y.; Zhang, Y.; Liu, W.; Huang, Y.; Shen, X.; Jing, R.; Pu, J.; Wang, X.; Ju, S.; Cong, H. LncRNA H19 overexpression induces bortezomib resistance in multiple myeloma by targeting MCL-1 via miR-29b-3p. Cell Death Dis. 2019, 10, 1–14. [Google Scholar] [CrossRef] [Green Version]
  249. Wu, Y.; Wang, H. LncRNA NEAT1 promotes dexamethasone resistance in multiple myeloma by targeting miR-193a/MCL1 pathway. J. Biochem. Mol. Toxicol. 2018, 32, e22008. [Google Scholar] [CrossRef]
  250. Tando, T.; Ishizaka, A.; Watanabe, H.; Ito, T.; Iida, S.; Haraguchi, T.; Mizutani, T.; Izumi, T.; Isobe, T.; Akiyama, T. Requiem protein links RelB/p52 and the Brm-type SWI/SNF complex in a noncanonical NF-κB pathway. J. Biol. Chem. 2010, 285, 21951–21960. [Google Scholar] [CrossRef] [Green Version]
  251. De Donatis, G.; Pape, E.; Pierron, A.; Cheli, Y.; Hofman, V.; Hofman, P.; Allegra, M.; Zahaf, K.; Bahadoran, P.; Rocchi, S. NF-kB2 induces senescence bypass in melanoma via a direct transcriptional activation of EZH2. Oncogene 2016, 35, 2735–2745. [Google Scholar] [CrossRef]
  252. Lee, J.-H.; Kang, M.-J.; Han, H.-Y.; Lee, M.-G.; Jeong, S.-I.; Ryu, B.-K.; Ha, T.-K.; Her, N.-G.; Han, J.; Park, S.J. Epigenetic alteration of PRKCDBP in colorectal cancers and its implication in tumor cell resistance to TNFα-induced apoptosis. Clin. Cancer Res. 2011, 17, 7551–7562. [Google Scholar] [CrossRef] [Green Version]
  253. van Essen, D.; Zhu, Y.; Saccani, S. A feed-forward circuit controlling inducible NF-κB target gene activation by promoter histone demethylation. Mol. Cell 2010, 39, 750–760. [Google Scholar] [CrossRef] [PubMed]
  254. Saccani, S.; Natoli, G. Dynamic changes in histone H3 Lys 9 methylation occurring at tightly regulated inducible inflammatory genes. Genes Dev. Biol. 2002, 16, 2219–2224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Hargreaves, D.C.; Horng, T.; Medzhitov, R. Control of inducible gene expression by signal-dependent transcriptional elongation. Cell 2009, 138, 129–145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Aberrant activation and transcriptional regulation of NF-κB pathway in cancer. In addition to constitutive activator molecules of both the canonical and non-canonical NF-κB pathway (highlighted in grey box), activation of the NF-κB pathway in cancer occurs via various other factors involving cytokines like TNF-α, IL-1, IL-17 secreted by tumour-associated macrophages (TAMs) and oncogenic driver mutations in various regulatory factors of the pathway (highlighted in pink box). Upon such activation, the activated NF-κB subunits interact with pro-tumorigenic transcription factors (highlighted in red oval), causing activation of alternative target genes associated with tumour promotion phenotypes, angiogenesis and tumour-associated non-coding RNAs.
Figure 1. Aberrant activation and transcriptional regulation of NF-κB pathway in cancer. In addition to constitutive activator molecules of both the canonical and non-canonical NF-κB pathway (highlighted in grey box), activation of the NF-κB pathway in cancer occurs via various other factors involving cytokines like TNF-α, IL-1, IL-17 secreted by tumour-associated macrophages (TAMs) and oncogenic driver mutations in various regulatory factors of the pathway (highlighted in pink box). Upon such activation, the activated NF-κB subunits interact with pro-tumorigenic transcription factors (highlighted in red oval), causing activation of alternative target genes associated with tumour promotion phenotypes, angiogenesis and tumour-associated non-coding RNAs.
Cells 12 00788 g001
Figure 2. Graphical summary of the mode of activation of oncogenic pathways by NF-κB signalling cascade via direct (bottom panel) and indirect (upper panel) interaction/regulations of various macromolecular components and factors.
Figure 2. Graphical summary of the mode of activation of oncogenic pathways by NF-κB signalling cascade via direct (bottom panel) and indirect (upper panel) interaction/regulations of various macromolecular components and factors.
Cells 12 00788 g002
Table 1. Summarized table of alteration of different components of NF-κB pathway in different cancer types (HCC: Human Colorectal Carcinoma; FLHCC: Fibrolamellar Hepatocellular Carcinoma; GBM: Glioblastoma Multiforme; MM: Multiple Myeloma; DLBCL: Diffuse large B-cell lymphoma; PMBL: Primary Mediastinal B-cell Lymphoma; ABC-DLBCL: Activated B-cell-like Diffuse large B-cell Lymphoma; PEL: Primary Effusion Lymphoma; ALT: Adult T-cell Lymphoma/Leukemia).
Table 1. Summarized table of alteration of different components of NF-κB pathway in different cancer types (HCC: Human Colorectal Carcinoma; FLHCC: Fibrolamellar Hepatocellular Carcinoma; GBM: Glioblastoma Multiforme; MM: Multiple Myeloma; DLBCL: Diffuse large B-cell lymphoma; PMBL: Primary Mediastinal B-cell Lymphoma; ABC-DLBCL: Activated B-cell-like Diffuse large B-cell Lymphoma; PEL: Primary Effusion Lymphoma; ALT: Adult T-cell Lymphoma/Leukemia).
Factors of NF-κB Pathway and AlterationsCancer TypeReferences
RelA (p65)
  • Higher expression and activation of p65 in tumour tissue compared to normal tissue
HCC, FLHCC, Breast cancer cell lines[105,114,115]
RelB
  • Frequent activation of RelB leading to non-canonical NF-κB pathway activation
GBM, MM, DLBCL[113,116,117,118,119]
c-Rel
  • Gain of chromosomal material (9p) leading enhanced expression of c-Rel
  • Gain of 2p leading to nuclear accumulation of c-Rel
Breast tumours, Hodgkin lymphoma, DLBCL, GC-DLBCL, PMBL[120,121,122]
NF-κB–inducing kinase (NIK)
  • Enhanced activation/overexpression of NIK in cancer cells/tumour compared to normal
GBM, MM, Hodgkin lymphoma[110,119,123]
IκBα
  • Enhanced degradation of IκBα causing endogenous activation of NF-κB pathway
  • High expression of IκBα transcript in certain cancer type
  • Direct transactivation via interaction with κB motifs in the DNA inside the nucleus
Ovarian Carcinoma cell line, HCC, Melanoma[104,105,109,124,125]
IKK
  • Higher expression of endogenous IKK in tumour tissue compared to normal tissue
  • Constitutive IKK activity causing enhanced activation of NF-κB pathway
Lung carcinoma, Melanoma, Hodgkin lymphoma, ABC-DLBCL, PEL, ALT, HCC[105,110,111,112,126,127,128]
NF-κB1
  • Higher expression of p50 in tumour tissue compared to normal tissue
  • Enhanced degradation of IκBα leads to increased activation of NF-κB1 subunit
Breast tumours, Melanoma cells, Lung cancer, HCC, ALT, ABC-DLBCL[104,105,110]
NF-κB2
  • Selective activation in some cancer types
  • Overexpression of NIK underlies constitutive activation of non canonical NF-κB pathway
Colon carcinoma cell line, MM, Breast cancer, Lung cancer[106,107,110,119,124]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Deka, K.; Li, Y. Transcriptional Regulation during Aberrant Activation of NF-κB Signalling in Cancer. Cells 2023, 12, 788. https://doi.org/10.3390/cells12050788

AMA Style

Deka K, Li Y. Transcriptional Regulation during Aberrant Activation of NF-κB Signalling in Cancer. Cells. 2023; 12(5):788. https://doi.org/10.3390/cells12050788

Chicago/Turabian Style

Deka, Kamalakshi, and Yinghui Li. 2023. "Transcriptional Regulation during Aberrant Activation of NF-κB Signalling in Cancer" Cells 12, no. 5: 788. https://doi.org/10.3390/cells12050788

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