Next Article in Journal
p38- and ERK-MAPK Signalling Modulate Developmental Neurotoxicity of Nickel and Vanadium in the Caenorhabditis elegans Model
Previous Article in Journal
The CK2/ECE1c Partnership: An Unveiled Pathway to Aggressiveness in Cancer
Previous Article in Special Issue
Interaction Networks Explain Holoenzyme Allostery in Protein Kinase A
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Yin and Yang of IκB Kinases in Cancer

by
Abdalla M. Abdrabou
1,2
1
Department of Biochemistry and Molecular Genetics, Feinberg School of Medicine, Northwestern University, Chicago, IL 60611, USA
2
Department of Medical Genetics, University of Alberta, Edmonton, AB T6G 2H7, Canada
Kinases Phosphatases 2024, 2(1), 9-27; https://doi.org/10.3390/kinasesphosphatases2010002
Submission received: 19 October 2023 / Revised: 22 December 2023 / Accepted: 29 December 2023 / Published: 31 December 2023
(This article belongs to the Special Issue Human Protein Kinases: Development of Small-Molecule Therapies)

Abstract

:
IκB kinases (IKKs), specifically IKKα and IKKβ, have long been recognized for their pivotal role in the NF-κB pathway, orchestrating immune and inflammatory responses. However, recent years have unveiled their dual role in cancer, where they can act as both promoters and suppressors of tumorigenesis. In addition, the interplay with pathways such as the MAPK and PI3K pathways underscores the complexity of IKK regulation and its multifaceted role in both inflammation and cancer. By exploring the molecular underpinnings of these processes, we can better comprehend the complex interplay between IKKs, tumor development, immune responses, and the development of more effective therapeutics. Ultimately, this review explores the dual role of IκB kinases in cancer, focusing on the impact of phosphorylation events and crosstalk with other signaling pathways, shedding light on their intricate regulation and multifaceted functions in both inflammation and cancer.

1. Introduction

The NF-κB (Nuclear Factor-κB) pathway is a linchpin of cellular responses to external stimuli, especially in the realms of immune and inflammatory processes [1]. Within the broader context of the NF-κB signaling pathway, the IκB kinases (IKKs) (Table 1), particularly IKKα and IKKβ, occupy a pivotal position in the intricate regulatory network of these pathways. Traditionally recognized for their roles in immune surveillance and defense, the IKKs have recently emerged as enigmatic figures in the landscape of cancer biology [2,3,4,5]. Their Janus-faced nature, promoting or suppressing tumorigenesis depending on context, has prompted an intensive exploration of their molecular mechanisms within the context of cancer and the possibility of targeting them [6].
The NF-κB pathway, with its intricate family members, represents a dynamic signaling network that orchestrates cellular responses to a multitude of extracellular signals [13]. The fundamental aspect of this pathway revolves around the NF-κB transcription factors, comprising various members such as p65 (RelA), RelB, c-Rel, p105/p50, and p100/p52. These members of the NF-κB family combine in different dimeric forms, each playing specific roles in overseeing signaling pathways, particularly those involved in immune responses [14,15,16].
Central to the regulation of NF-κB is the presence of inhibitory proteins referred to as IκBs, responsible for maintaining NF-κB dimers in an inactive state within the cell’s cytoplasm [17]. The activation of the pathway involves the phosphorylation of IκBs, tagging them for degradation by the proteasome machinery. This process liberates NF-κB dimers, allowing them to migrate into the nucleus and commence gene transcription (Figure 1) [18].
The NF-κB pathway, essential for various physiological processes, exists in multiple branches. Primarily, there are two well-characterized pathways: the canonical (or classical) and non-canonical (or alternative) [19]. These pathways differ in terms of the stimuli that activate them, the proteins involved, and the nature of their functions. The effective enhancement of NF-κB involves a complex regulation process mediated by the IκB kinase (IKK) complex. This intricate system orchestrates the phosphorylation of IκB proteins, leading to their ubiquitination and subsequent degradation via the proteasome [20,21,22]. This series of events ultimately results in the liberated NF-κB complexes translocating into the nucleus.
Within the nucleus, these NF-κB complexes engage with specific DNA sequences, thereby governing the transcription of genes involved in diverse processes, including immune responses, cellular growth regulation, and the modulation of cell survival [23,24,25]. Notably, within the context of cancer, NF-κB-dependent genes encompass those responsible for encoding cytokines, chemokines, cyclin D1, matrix metalloproteinases, and antiapoptotic proteins such as Bcl-xL [26,27,28].

2. Canonical NF-κB Pathway

In the Canonical Nuclear Factor-kappa B (NF-κB) pathway, activation is initiated by a diverse array of stimuli, encompassing proinflammatory cytokines like Tumor Necrosis Factor-alpha (TNF-α) and Interleukin-1 beta (IL-1β), as well as microbial products like lipopolysaccharides (LPSs) (Figure 1) [29,30,31,32]. Key to this pathway is the activation of IKKβ, which subsequently phosphorylates and targets IκBα and IκBβ for degradation. As a result, this liberation allows for the movement of p50-RelA dimers into the nucleus. Once in the nucleus, these dimers function as transcription factors, overseeing the expression of genes linked to inflammation, immune responses, and cell survival (Figure 1) [33,34,35].

3. Non-Canonical NF-κB Pathway

Conversely, the Non-Canonical NF-κB pathway is typically triggered by a unique group of receptors, which encompass the lymphotoxin-β receptor (LTβR) and B-cell activating factor receptor (BAFF-R) [36]. This pathway is reliant on the conversion of p100 to p52, a process orchestrated by the activation of IKKα. Subsequently, the p52-RelB dimers relocate to the nucleus, assuming critical roles in the development of secondary lymphoid organs, B-cell maturation, and the organization of lymphoid tissues (Figure 1) [37,38,39].
NIK, or NF-κB-inducing kinase, holds a crucial position in regulating the non-canonical NF-κB signaling pathway [40]. Initially acknowledged for its role in activating the canonical NF-κB pathway, the absence of NIK did not hinder the TNF-induced IKKβ/p65/p50 activation. However, it was later discovered to be essential for triggering the non-canonical NF-κB pathway [41,42,43]. The regulation of NIK predominantly occurs post-translationally. Structurally, NIK encompasses four domains: a TRAF3-binding N-terminal region, a negative regulatory domain (NRD), a core kinase domain, and a C-terminal domain responsible for binding with proteins like IKKα and p100 [44,45,46].
Initially recognized as a mediator following TNF and IL-1 receptor activation, NIK’s kinase activity was deemed crucial in facilitating this particular process. Additionally, it mediates stimulation through various receptors like CD27, CD30, CD40, LTβR, and BAFFR [47,48,49,50]. The overexpression of NIK activates NF-κB, protecting cells from TNF-induced apoptosis, while kinase-dead NIK mutants inhibit NF-κB activation by TNFα [51,52,53].
Under normal conditions, NIK binds to TRAF2/3 and cIAP1/2, leading to its continuous ubiquitination and degradation [54]. Stimulation by cytokines (such as CD40L, TWEAK, LTα/β, or LPS) sequesters TRAF2/3, allowing the cIAP1-mediated ubiquitination of TRAF3 [55]. The subsequent degradation of TRAF3 leads to the accumulation of newly synthesized NIK within the cell. This stabilization and buildup of NIK are crucial for initiating the noncanonical NF-κB pathway [56,57,58]. Upon receptor activation, NIK triggers IKKα phosphorylation at Ser-176 and Ser-180, activating it to phosphorylate p100. The phosphorylation of p100 prompts the binding to ubiquitin ligase β-TrCP, resulting in partial proteasomal processing to p52. This processing removes the inhibitory C-terminal ankyrin repeat domain of p100, akin to the function of mature IκB proteins, thus maintaining RelB inactive in the cytoplasm. Subsequently, p52-RelB translocates to the nucleus to regulate transcription [59,60,61,62].
NIK interacts with and activates both IKKα and IKKβ, phosphorylating IKKα to a greater extent. Consequently, NIK acts as an upstream kinase for the IKK complex, facilitating signaling from multiple cytokine receptors [63,64,65]. Several other kinases, including MEKK1 and TAK1, were identified as IKK kinases, sometimes acting alongside NIK. TAK1, for instance, can activate NIK/IKK/NF-κB signaling independently of NIK in certain contexts [66]. Additionally, proteins like TRAF2, 5, and 6, as well as TBK-1, contribute to NF-κB activation by acting upstream of NIK. Moreover, Bcl10 has been reported to phosphorylate NIK under specific inflammatory conditions in human colonic epithelial cells treated with carrageenan (CGN) [22,67,68].

4. The Dark Side: IκB Kinases as Tumor Promoters

4.1. IKKα (Inhibitor of κB Kinase Alpha)

IKKα plays a multifaceted role in cancer, impacting both its initiation and progression, along with metastasis. In colorectal cancer cells (HT29), IKKα exhibits abnormal activation within the nucleus of tumor cells [69]. Here, it binds to specific genes reliant on Notch signaling, such as hes1 and herp2. The nuclear IKKα phosphorylates a nuclear co-repressor, SMRT, causing its release from chromatin and the subsequent expression of Notch-dependent genes [70], leading to more aggressive growth and proliferation. Pan-IKK inhibition re-establishes SMRT chromatin binding, curbing Notch-related gene expression, and restraining tumor growth in experimental models [71].
Additionally, IKKα phosphorylates N-CoR, akin to SMRT, facilitating its nuclear export from CRC cells [72]. The active nuclear IKKα isoform, IKKα(p45), is crucial for preventing apoptosis and thereby fostering tumor growth, specifically in HCT116 cells. Mechanistically, the association between active TAK1, BRAF, a complex containing IKKα(p45), and NEMO leads to SMRT and Histone H3 phosphorylation, which is vital for BRAF-mediated transformation independent from NF-κB signaling [73].
In keratinocytes, evidence demonstrates IKKα’s involvement in cancer initiation independently of NF-κB [74]. The deletion of IKKα induces skin squamous cell carcinoma in mice, affecting 14-3-3σ expression and prompting aberrant cell proliferation, disrupting skin homeostasis, and promoting cell transformation [75]. Additional studies support IKKα’s tumor suppressor role in the skin, linking its activity to the transforming growth factor beta (TGFβ) pathway. Moreover, a specific variant of nuclear IKKα in keratinocytes leads to more aggressive tumors upon exposure to chemical carcinogens [76,77].
Basal cell carcinomas (BCCs) are the most prevalent among human cancers affecting the skin [78]. While the noncanonical NF-κB pathway relies on IKKα, its specific role in BCC remains unclear. One study indicated that, within both BCC and non-malignant conditions, IKKα is present in the nucleus. Within BCC, the nuclear IKKα directly interacts with the promoters of inflammation factors and LGR5, a marker for stem cells. This interaction leads to an increase in LGR5 expression through the activation of the STAT3 signaling pathway, thereby contributing to cancer progression. The activation of the STAT3 pathway influences the LGR5 expression in a manner dependent on IKKα, as demonstrated by the interplay between STAT3 and IKKα. Moreover, suppressing the IKKα impedes the tumor growth and transition from the epithelial stage to the mesenchymal stage. This finding highlights IKKα’s role as a genuine chromatin regulator in BCC. Its heightened expression facilitates oncogenic transformation by promoting the expression of genes related to stemness and inflammation. Consequently, these findings offer a fresh perspective on how IKKα may participate in the progression of BCC tumors within an inflammatory microenvironment [79].
Moreover, in a study by Mahato and colleagues [80], they have shown that the suppression of IKKα in prostate cancer cells using synthetic siRNAs affects tumor cell growth and invasiveness. In this study, the authors designed three synthetic siRNAs targeting the specific regions of IKKα mRNA and evaluated their ability to silence IKKα in PC-3 and DU145 cells. A range of assays, including wound healing, migration, proliferation, and cell cycle analysis, were employed to investigate how IKKα siRNAs biologically impacted prostate cancer cells. Interestingly, their results uncovered potent siRNAs that could silence IKKα by up to 70%, resulting in decreased wound healing, migration, invasion, and cell attachment capabilities in prostate cancer cells. Additionally, this study observed comparable anti-invasive effects in the presence of RANKL. However, silencing IKKα had minimal effects on cell proliferation and cell cycle distribution. These findings strongly indicate that IKKα significantly influences prostate cancer invasion and metastasis while playing a minor role in cell proliferation. Targeting IKKα using siRNA emerges as a promising therapeutic approach for managing prostate cancer by reducing invasion and metastasis without directly impacting cell proliferation [81,82].
Moreover, IKKα contributes to progesterone-induced tumor promotion in breast cancer, downstream of RANKL induction, and fosters metastatic spread relying on RANKL produced by tumor-infiltrating regulatory T cells. It phosphorylates Estrogen Receptor α, its coactivator AIB1/SRC3, and induces targets like cyclin D1 and c-myc, driving breast cancer cell proliferation [83]. Clinical observations link IKKα expression in breast cancer cells with patient outcomes regardless of cellular localization. In triple-negative breast cancer (TNBC) cells, IKKα mediates Notch signaling triggered by the Notch ligand Jagged1, a pivotal pathway for TNBC Cancer Stem Cell survival [84]. Combining therapies targeting the intersection of Notch, AKT, and NF-κB pathways holds promise for therapeutic applications against cancer stem cells in TNBC [85].

4.2. IKKβ (Inhibitor of κB Kinase Beta)

In contrast to IKKα, IKKβ is predominantly associated with the canonical NF-κB pathway. One of its key functions is the phosphorylation of IκBα and IκBβ. This phosphorylation event marks IκB proteins for degradation, allowing the release of p50-RelA dimers [86]. The degradation of IκB proteins leads to the liberation of NF-κB, a transcription factor crucial in orchestrating the expression of pro-inflammatory genes. This includes genes responsible for cytokine and chemokine production, thus fostering a sustained and amplified inflammatory response within the tumor microenvironment [87]. This persistent inflammation, driven by IKKβ, creates a milieu that nurtures tumor growth, angiogenesis, and metastasis [88].
The released dimers subsequently move into the nucleus, where they commence the transcription of genes linked with the canonical NF-κB pathway. Notably, the activation of IKKβ is frequently prompted by proinflammatory stimuli, connecting it to persistent inflammation [89]. This linkage further underscores IKKβ’s significance in fostering tumor progression, the formation of new blood vessels (angiogenesis), and the spread of cancer to distant sites (metastasis), highlighting its critical role in the context of cancer. However, it is important to note that the outcome of this inflammatory response is context-dependent, either influencing the promotion of tumor formation or the initiation of an immune reaction against tumors [90].
To date, diverse chemical inhibitors targeting IKKβ have been discovered, each employing distinct mechanisms [91] (see Table 2). Most of these inhibitors mimic ATP, displaying reversible, ATP-competitive behavior, often exhibiting some preference for inhibiting IKKβ over IKKα and other kinases [92]. Yet, due to the structural similarity of protein kinase ATP-binding sites, these ATP mimics can inadvertently affect other kinases, causing unintended effects at concentrations required to inhibit their primary target in cells [93]. Specifically, some commonly used ‘specific’ IKKβ inhibitors, such as Bay 11-7082 and TPCA-1, have been found to induce significant off-target effects [94]. For instance, Bay 11-7082 disrupts NF-κB by irreversibly deactivating the E2-conjugating enzymes Ubc13 and UbcH7, as well as the E3-ligase LUBAC, rather than directly inhibiting IKK activity. Similarly, TPCA-1 hampers the STAT3 signaling by directly binding to the STAT3 Src Homology 2 (SH2) domain, alongside its IKKβ inhibitory activity [95].
Currently, the most potent ATP-competitive inhibitors for IKKβ include MLN-120B and BI605906, showcasing over 50-fold and over 300-fold selectivity for IKKβ over IKKα, respectively [105].
Recent research indicates potential toxicity and side effects correlated with IKKβ inhibition, such as the onset of inflammatory skin diseases and the heightened vulnerability of colonic epithelium to various stressors [106,107]. Severe liver malfunction has been observed in mice with IKKβ deficiencies, and intestinal and liver toxicity has surfaced in numerous clinical trials involving IKKβ inhibitors, potentially restricting their clinical applicability [108].

5. The Bright Side: IKKs as Tumor Suppressors

5.1. IKKα in Tumor Suppression

Despite its role in promoting cancer in some contexts, IKKα also has tumor-suppressive functions. It can induce cellular senescence in response to oncogenic stress, causing cells to enter a state of irreversible growth arrest [109]. Cells that undergo senescence not only cease dividing but also release substances referred to as the senescence-associated secretory phenotype (SASP). These substances attract immune cells, which in turn play a role in eliminating cells that could potentially develop into tumors [110].
A significant association exists between DNA damage-triggered senescence and the NF-κB-regulated SASP [111]. When exposed to genotoxic stress, the Ataxia telangiectasia mutated (ATM) kinase activates NF-κB by triggering the post-translational modifications (PTMs) of NEMO, which play a pivotal role in NF-κB activation. NEMO activation by ATM subsequently triggers the IKK complex, culminating in the nuclear translocation of NF-κB and the transcription of numerous genes related to SASP [112]. In melanoma, senescent cells produce a secretome characterized by the pro-invasive and pro-tumorigenic properties, relying on PARP-1 and NF-κB [113].
The expression of SASP components IL-6 and IL-8 necessitates IκBζ during both DNA damage-induced senescence and oncogene-induced senescence (OIS), establishing IκBζ as a crucial modulator of the proinflammatory SASP [114].
Multiple strands of evidence indicate that the continuous DNA damage response (DDR) is crucial for robust SASP production. The depletion of DDR elements like ATM, NBS1, or CHK2 inhibits the expression of IL-6, IL-8, and several GRO family members [115]. Consequently, it has been demonstrated, at least in specific experimental setups, that DDR activation—not just the presence of DNA damage itself—governs senescent states and SASP regulation [116].
Metformin, an anti-diabetic medication with diverse effects, also exerts activity on senescent cells [117]. One of its effects involves impairing the SASP of RAS-induced senescent cells without impeding proliferative arrest. This occurs through the inhibition of IKKα/β and IκB phosphorylation by metformin, preventing the nuclear translocation of p65 (RelA) [118]. Metformin negatively influences NF-κB without affecting other inflammatory pathways like p38, JNK, and IRF. The inhibition of SASP by metformin might contribute to the observed anti-aging effects post-metformin treatment [119].

5.2. IKKβ in Antitumor Immune Responses

One of the central mechanisms through which IKKβ contributes to antitumor immunity is the activation of immune cells, particularly T cells and dendritic cells (DCs) [2]. These immune cells play pivotal roles in orchestrating immune responses against cancer.
IKKβ activation in T cells enhances their responsiveness and effector functions. T cells are the foot soldiers of the immune system, responsible for recognizing and eliminating cancer cells [120]. The activation of the IKKβ/NF-κB pathway in T cells augments their activation and proliferation. This, in turn, leads to an increased pool of cytotoxic T lymphocytes (CTLs) that can effectively target and kill cancer cells [121]. Additionally, activated T cells can infiltrate the tumor microenvironment, exerting their antitumor effects directly at the site of malignancy.
T cells capable of recognizing tumor-associated antigens exhibit potential in eradicating tumors [122]. Despite their presence in cancer patients—both in circulation and within tumors—these tumor-reactive T cells often fail to prevent tumor progression over time, indicating a probable decline in their functional abilities [123]. The direct analysis of these tumor antigen-specific T cells revealed deficiencies in cytokine production and cytolytic activity. Efforts to intervene and restore T cell function have shown promise in clinical settings but frequently result in partial responses. Understanding the mechanisms underlying T cell dysfunction in cancer remains crucial for enhancing therapeutic efficacy [124].
One critical pathway for T cell function involves the activation of IκB kinase β (IKKβ) and the subsequent activation of NF-κB. In the tumor environment, T cell-NF-κB activity is often hindered, leading to reduced functionality in T cells isolated from cancer patients [125]. Recent studies conducted with mouse models exhibiting compromised NF-κB downstream of the T cell receptor (TCR) underscored the critical importance of T cell-NF-κB activation in the release of cytokines, specific targeting, and the destruction of antigens, and the in vivo eradication of tumors [126]. This indicates that reduced T cell-NF-κB activity induced by growing tumors compromises anti-tumor T cell responses, fostering a cycle favoring tumor growth. Consequently, exploring methods to stimulate T cell-intrinsic NF-κB activity becomes a compelling avenue for enhancing anti-tumor immunity [127,128].
In a study by Evaristo et al. [129], novel genetic mouse models expressing constitutively active IKKβ (caIKKβ) specifically in T cells were employed. The results demonstrated that the T cell-specific expression of caIKKβ significantly improved tumor control, even in cases of established tumors. Thus, stimulating T cell-intrinsic NF-κB appears crucial in responding to cancer growth, suggesting that the therapeutic manipulation of the IKKβ/NF-κB axis holds promise for boosting anti-tumor immune responses.
DCs are antigen-presenting cells that play a critical role in initiating and shaping antitumor immune responses. IKKβ activation in DCs enhances their ability to capture, process, and present tumor antigens to T cells [130]. This process, known as antigen presentation, is a crucial step in initiating an adaptive immune response against cancer. The activation of the IKKβ/NF-κB pathway in DCs results in the upregulated expression of co-stimulatory molecules and cytokines that are necessary for efficient T-cell priming [131].
IKKβ activation in DCs leads to the upregulation of co-stimulatory molecules, such as CD80 and CD86 [132]. These molecules interact with their corresponding receptors on T cells, providing essential co-stimulatory signals that are required for T-cell activation and proliferation. The enhanced expression of these co-stimulatory molecules by IKKβ-activated DCs amplifies the effectiveness of T-cell priming and the subsequent antitumor immune response [133].
Baratin et al., 2015 [134] conducted a comparative analysis of the transcriptomes of NLT-DCs within the skin and their migratory counterparts located in the draining lymph nodes (LNs). Through this investigation, they identified a novel gene network that is regulated by the NF-kB pathway and is specific to migratory dendritic cells. Their findings demonstrate that the targeted deletion of IKKβ, a key activator of NF-kB, in dendritic cells, hampers the accumulation of NLT-DCs in LNs and impairs the conversion of regulatory T cells in vivo. These outcomes are closely associated with disruptions in immune tolerance and the onset of autoimmune responses.
The activation of IKKβ leads to a heightened production of proinflammatory cytokines like interleukin-12 (IL-12) and interferon-gamma (IFN-γ) by DCs [135]. These cytokines serve a pivotal function in fostering an immune environment that supports anti-tumor immunity. IL-12, for instance, can skew the immune response towards a Th1 phenotype, characterized by enhanced cytotoxic activity and IFN-γ production by T cells. IFN-γ, on the other hand, has direct antitumor effects and can activate other immune cells to contribute to tumor eradication [136].

6. Distinct Phosphorylation Events and Kinases

Serine/threonine phosphorylation by IKKα and IKKβ plays a pivotal role in regulating the NF-κB pathway. When NF-κB is inactive, it is typically sequestered in the cytoplasm by inhibitor proteins known as IκB (Inhibitor of κB) [137]. The phosphorylation of serine residues within IκB proteins, especially IκBα and IκBβ, is a critical event orchestrated by IKKβ. This phosphorylation serves as a recognition signal for the E3 ubiquitin ligase, which ubiquitinates IκBα and IκBβe [138].
Tyrosine-phosphorylated STAT dimers represent the culmination of these intricate processes. Once formed, these dimers are ready to exert their transcriptional influence. STAT dimers translocate from the cytoplasm into the cell nucleus. Inside the nucleus, they bind to specific DNA sequences known as enhancer elements or response elements in the regulatory regions of target genes [139]. This binding event is specific to the dimer’s composition and the cytokine signals received.
The binding of tyrosine-phosphorylated STAT dimers to these regulatory elements serves as a molecular switch, initiating the transcription of genes associated with immune responses and inflammatory processes [140]. The regulated genes often encode critical immune effectors, signaling molecules, and cytokines, shaping the cell’s response and ultimately contributing to the immune and inflammatory outcomes observed in response to cytokine stimulation. Importantly, the activation of tyrosine kinases, such as JAKs, indirectly stimulates the NF-κB pathway [141]. This occurs through the cooperative efforts of transcription factors like STATs, which activate the gene expression related to immune responses. These genes may include those encoding proinflammatory cytokines and chemokines [142].
The coordination of various signaling pathways, involving both tyrosine phosphorylation and serine/threonine phosphorylation events, ensures that the cellular response is robust, efficient, and finely tuned to the needs of the immune system [143]. For instance, in viral infection, the release of interferons triggers the activation of JAK–STAT signaling, leading to the transcription of antiviral genes [144]. At the same time, the activation of the NF-κB pathway stimulates the production of proinflammatory cytokines, which, in turn, serve to attract immune cells to the site of infection. The interplay between these signaling pathways optimizes the host’s response to the viral threat, underscoring the complexity of cross-talk mechanisms in immune regulation [145].
In the non-canonical NF-κB activation pathway, IKKα primarily phosphorylates a different substrate, p100. This phosphorylation leads to a unique processing event that is crucial for the activation of non-canonical NF-κB [146].
The phosphorylation of p100 by IKKα is followed by its partial proteasomal processing. This processing event results in the generation of a smaller protein fragment, p52. Importantly, p52 contains the DNA-binding domain necessary for transcriptional activity, allowing it to function as an NF-κB transcription factor [147]. The p52 subunit combines with other proteins, like RelB, to create dimers that then move into the nucleus. This sequence of events triggers the activation of the non-canonical NF-κB pathway, which functions differently compared to the canonical pathway [148].
The mechanism behind the phosphorylation of the IκB kinase T-loop remains a significant unanswered query. Suggestions have arisen proposing the involvement of IKK kinases (IKKKs) in this process, drawing an analogy to other signaling pathways [149]. One prominent example is TAK1, known for its involvement in the JNK pathway [150]. In cell-free assays, TAK1, along with the adaptor proteins TAB 1 and TAB 2, has been identified as a TRAF6-regulated IKK activator [151]. TAB 2’s role in recruiting TAK1 to the K63-linked polyubiquitin chains of upstream regulators likely induces IKKβ phosphorylation through proximity-driven mechanisms [152]. However, TAK1 is not a universal IKKK but rather functions as a regulatory module impacting the IKK activation based on stimuli and cell type. Another potential IKKK, MEKK3 [153], has been proposed as it can phosphorylate IKK in vitro, and its deficiency correlates with the reduced NF-κB activation in response to various stimulations like TNF, IL-1, or TLR. IL-1-induced NF-κB activation has been linked to MEKK3 alongside TAK1 [154]. However, it is also suggested that IKK subunits might undergo activation through trans-autophosphorylation rather than via an IKKK. Recent structural and composition analyses support these possibilities, indicating that trans-autophosphorylation and IKKK-dependent phosphorylation might work sequentially or in parallel to achieve optimal kinase activation [155].
The dimeric NF-κB transcription factor, consisting of Rel family subunits that bind to DNA, plays a pivotal role in immune and inflammatory responses. NF-κB has recently been identified as a protector against apoptosis induced by tumor necrosis factor (TNF) and various genotoxic agents [156]. Typically, NF-κB dimers are confined within the cytoplasm due to their interaction with inhibitory IκB proteins. These proteins bind to the Rel homology domain (RHD), which is responsible for the dimerization, nuclear translocation, and DNA binding functions of NF-κB/Rel proteins [157]. When cells are stimulated by proinflammatory cytokines (e.g., IL-1, TNF), bacterial lipopolysaccharide (LPS) or phorbol ester (TPA), specific IκBs (such as IκBα and IκBβ) undergo rapid phosphorylation at certain N-terminal residues. This phosphorylation leads to their subsequent polyubiquitination and degradation via the 26S proteasome [158].
The breakdown of these inhibitory IκB proteins liberates the NF-κB dimer, enabling its migration into the nucleus to commence gene transcription [159]. Altering the phosphorylation sites on IκBα through specific mutations prevents its phosphorylation, ubiquitination, and subsequent degradation. Mutants like IκBα(A32/36) act as powerful inhibitors of NF-κB activation. Efforts to comprehend the regulation of this pathway have concentrated on identifying the responsible protein kinase(s). The IκB kinase (IKK) complex, a 900 kDa protein kinase complex, phosphorylates IκBα and IκBβ at sites crucial for their ubiquitination and degradation [160]. Another protein kinase complex phosphorylating IκB has been described, but its relationship with IKK remains unclear. IKK’s activity is rapidly stimulated by IL-1 or TNF and is dependent on its phosphorylation [161]. One component of the IKK complex, IKKα, an 85 kDa polypeptide containing a protein kinase domain and protein interaction motifs, has been molecularly identified. IKKα’s expression is vital for NF-κB activation by various stimuli. IKKα has been isolated as a NIK-interacting protein, suggesting its role in NIK-mediated NF-κB activation [162]. While a catalytically inactive IKKα mutant can inhibit NF-κB activation, it can still interact with other IKK components to form a functional but less active IκB kinase complex.
Ubiquitination is a post-translational modification that tags proteins for degradation by the proteasome [163]. Following ubiquitination, IκBα and IκBβ are targeted for proteasomal degradation. As a result, they are rapidly degraded, freeing NF-κB from its sequestration. The released NF-κB transcription factors can subsequently migrate to the cell nucleus and initiate the transcription of genes [164].
The activation of IKKs appears to hinge on induced proximity from densely organized signaling complexes and the binding of adapter proteins like NEMO or TAB proteins. Non-degradative polyubiquitination, essential for IKK complex activation, triggers these processes [165].
TRAF6 stands as the initial ubiquitin E3 ligase identified, catalyzing K63-linked auto-ubiquitination alongside Ubc13 and Uev1A, subsequently initiating IKK activation. TRAF6 participates in numerous NF-κB-stimulating signaling pathways, including those triggered by IL-1R, TLR, TCR, RIG-I-like receptor, and DNA double-strand breaks [166]. In IL-1 signaling, TRAF6’s enzymatic activity, rather than auto-ubiquitination, is essential for NF-κB activation. TRAF6 not only self-ubiquitinates but also mediates the K63-linked polyubiquitination of various pathway components like IRAK1, MALT1, and TAK1 [167]. Moreover, it is suggested that TRAF6 generates free, unanchored K63-linked polyubiquitin, serving as a docking platform in IKK activation [168]. Several K63-specific E3 ligases involved in distinct NF-κB signaling cascades were identified, such as TRAF2/5, or TRIM25 in the TNFR, IL-1R/TLR, and RIG-I pathways. Additionally, several proposed NF-κB pathway regulators are substrates of inducible K63 ubiquitination, including Bcl10, NOD2, and ELKS [169].
TNFα signaling does not rely on K63-linked ubiquitination, suggesting the importance of alternative non-degradative polyubiquitination in this pathway [170]. For instance, the linear, M1-linked ubiquitination of NEMO and RIP1 by the LUBAC complex is crucial for NF-κB activation. LUBAC is associated with the TNFR1 signaling complex. LUBAC-mediated M1-linked ubiquitination contributes to various NF-κB activations but is dispensable for B-cell receptor signaling. Mutations affecting LUBAC components were linked to immune-related disorders, showcasing their physiological relevance [171].
Multiple ubiquitin linkages seem to play roles in NF-κB signaling, adding to the complexity of ubiquitin-mediated processes [172]. E3 ligases like cIAP1 and TRIM23 catalyze distinct ubiquitin chain types, and certain proteins show modifications with various ubiquitin linkages in response to stimuli. Additionally, the mono-ubiquitination of proteins like NEMO has been shown to impact NF-κB activation. These diverse modifications coordinate specific protein interactions in NF-κB signaling pathways, although many intricacies regarding these actions remain undiscovered [173].

7. Genetic and Epigenetic Regulation

While primary genetic mutations within the IκB kinase genes themselves are relatively rare occurrences in the realm of cancer, modifications affecting their upstream regulatory elements can exert profound and far-reaching effects. Somatic mutations affecting lysine 171 within the IKBKB gene, responsible for encoding the critical activating kinase (IKKβ) in the canonical NFκB signaling pathway, were observed in splenic marginal zone lymphomas and multiple myeloma. Lysine 171 is part of a positively charged pocket crucial for interaction with the activation loop phosphate in the naturally activated kinase [174].
Their findings demonstrate that both K171E IKKβ and K171T IKKβ variants function as kinases that remain constantly active, even in the absence of activation loop phosphorylation. Through predictive modeling and biochemical investigations, we elucidate why mutations in a positively charged residue within the cationic pocket of a kinase reliant on activation loop phosphorylation lead to persistent activation [175].
Utilizing transcription activator-like effector nuclease (TALEN)-based knock-in mutagenesis, we present evidence from a B lymphoid context indicating the involvement of K171E IKKβ in the development of lymphomas. Genetic mutations in TRAF proteins have the potential to engender the persistent and aberrant activation of the NF-κB pathway, a phenomenon observed in various cancer types [176]. A notable instance is the frequent occurrence of mutations in the TRAF3 gene within the context of multiple myeloma, which leads to a chronic state of NF-κB pathway activation. The discernment of such genetic anomalies, often made possible through advanced techniques like whole-genome sequencing, not only advances our comprehension of the multifaceted terrain of cancer biology but also unveils promising therapeutic targets for potential intervention [177].
Multiple mechanisms contribute to NF-κB transcriptional regulation beyond its binding to κB regulatory elements in DNA. In non-neuronal cells, the signaling elements of the NF-κB pathway participate in gene expression control through histone phosphorylation and acetylation in coordination with histone deacetylases (HDACs). The IκB protein variant, IκBα, independently governs transcription by engaging with HDAC1 and HDAC3. Additionally, the IKKα subunit operates distinctly from the IKK complex, influencing the cytokine-induced gene expression by modulating histone H3 phosphorylation [178]. These investigations reveal the novel functions of NF-κB signaling components, like IκBα and IKKα, in autonomously regulating the chromatin structure and gene expression from NF-κB’s direct DNA binding.
Previously formed memories are prone to disruption immediately post-recall, necessitating reconsolidation. While protein translation mechanisms are acknowledged for their role in memory reconsolidation, research into gene transcription mechanisms remains relatively limited in this context [179].
An interesting study [180] indicated that the retrieval of contextual conditioned fear memories activates the NF-κB pathway, regulating histone H3 phosphorylation and acetylation at specific gene promoters in the hippocampus, particularly mediated by IKKα rather than the NF-κB DNA-binding complex. Behaviorally, inhibiting IKKα’s control over a chromatin structure or NF-κB DNA-binding complex activity results in impairments in fear memory reconsolidation. Elevated histone acetylation offsets this memory deficit when faced with the IKK blockade. These results offer new insights into IKK-mediated transcriptional mechanisms in the hippocampus essential for memory reconsolidation.
MicroRNAs (miRNAs) are another essential post-transcriptional regulator of IκB kinases [181]. One example is miR-21, a well-known oncogenic miRNA that plays a role in regulating the NF-κB pathway in various cancers. In certain scenarios, miR-21 targets PTEN (Phosphatase and Tensin Homolog), an inhibitor of the NF-κB pathway. This targeting leads to increased NF-κB activity in cancer cells, which, in turn, promotes cell survival, proliferation, and resistance to apoptosis [182].

8. Conclusions

IκB kinases, IKKα and IKKβ, represent a multifaceted and pivotal aspect of cancer biology, serving as central players in inflammation, immune responses, and tumorigenesis. Their dual role in cancer, which is shaped by myriad factors including distinct phosphorylation events, genetic and epigenetic regulation, and therapeutic implications, underscores the intricacies of their functions.
The potential of IκB kinases as therapeutic targets in cancer therapy is a burgeoning field. Precision medicine, which aims to individualize the treatment strategies based on genetic and molecular profiles, is at the forefront of this effort. While challenges such as resistance to IKK inhibitors persist, the development of combinatorial therapies and immunomodulation strategies offers hope for overcoming these obstacles.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mitchell, S.; Vargas, J.; Hoffmann, A. Signaling via the NFκB system. Wiley Interdiscip. Rev. Syst. Biol. Med. 2016, 8, 227–241. [Google Scholar] [CrossRef] [PubMed]
  2. Mulero, M.C.; Huxford, T.; Ghosh, G. NF-κB, IκB, and IKK: Integral Components of Immune System Signaling. Adv. Exp. Med. Biol. 2019, 1172, 207–226. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, F.; Xia, Y.; Parker, A.S.; Verma, I.M. IKK biology. Immunol. Rev. 2012, 246, 239–253. [Google Scholar] [CrossRef] [PubMed]
  4. Durand, J.K.; Baldwin, A.S. Targeting IKK and NF-κB for Therapy. Adv. Protein Chem. Struct. Biol. 2017, 107, 77–115. [Google Scholar] [CrossRef] [PubMed]
  5. Antonia, R.J.; Hagan, R.S.; Baldwin, A.S. Expanding the View of IKK: New Substrates and New Biology. Trends Cell Biol. 2021, 31, 166–178. [Google Scholar] [CrossRef]
  6. Senegas, A.; Gautheron, J.; Maurin, A.G.; Courtois, G. IKK-related genetic diseases: Probing NF-κB functions in humans and other matters. Cell. Mol. Life Sci. 2015, 72, 1275–1287, Erratum in Cell. Mol. Life Sci. 2015, 72, 1289. [Google Scholar] [CrossRef] [PubMed]
  7. Huang, W.C.; Hung, M.C. Beyond NF-κB activation: Nuclear functions of IκB kinase α. J. Biomed. Sci. 2013, 20, 3. [Google Scholar] [CrossRef]
  8. Gu, L.; Zhu, Y.; Lin, X.; Lu, B.; Zhou, X.; Zhou, F.; Zhao, Q.; Prochownik, E.V.; Li, Y. The IKKβ-USP30-ACLY Axis Controls Lipogenesis and Tumorigenesis. Hepatology 2021, 73, 160–174. [Google Scholar] [CrossRef]
  9. Al Hamrashdi, M.; Brady, G. Regulation of IRF3 activation in human antiviral signaling pathways. Biochem. Pharmacol. 2022, 200, 115026. [Google Scholar] [CrossRef]
  10. Fitzgerald, K.A.; McWhirter, S.M.; Faia, K.L.; Rowe, D.C.; Latz, E.; Golenbock, D.T.; Coyle, A.J.; Liao, S.M.; Maniatis, T. IKKepsilon and TBK1 are essential components of the IRF3 signaling pathway. Nat. Immunol. 2003, 4, 491–496. [Google Scholar] [CrossRef]
  11. Willems, M.; Dubois, N.; Musumeci, L.; Bours, V.; Robe, P.A. IκBζ: An emerging player in cancer. Oncotarget 2016, 7, 66310–66322. [Google Scholar] [CrossRef] [PubMed]
  12. Yu, Z.; Gao, J.; Zhang, X.; Peng, Y.; Wei, W.; Xu, J.; Li, Z.; Wang, C.; Zhou, M.; Tian, X.; et al. Characterization of a small-molecule inhibitor targeting NEMO/IKKβ to suppress colorectal cancer growth. Signal Transduct. Target. Ther. 2022, 7, 71, Erratum in Signal Transduct. Target. Ther. 2023, 8, 330. [Google Scholar] [CrossRef] [PubMed]
  13. Zhu, F.; Hu, Y. Integrity of IKK/NF-κB Shields Thymic Stroma That Suppresses Susceptibility to Autoimmunity, Fungal Infection, and Carcinogenesis. Bioessays 2018, 40, e1700131. [Google Scholar] [CrossRef] [PubMed]
  14. Courtois, G.; Israël, A. IKK regulation and human genetics. Curr. Top. Microbiol. Immunol. 2011, 349, 73–95. [Google Scholar] [CrossRef] [PubMed]
  15. May, M.J.; Ghosh, S. Signal transduction through NF-κB. Immunol. Today 1998, 19, 80–88. [Google Scholar] [CrossRef] [PubMed]
  16. Hellweg, C.E. The Nuclear Factor κB pathway: A link to the immune system in the radiation response. Cancer Lett. 2015, 368, 275–289. [Google Scholar] [CrossRef] [PubMed]
  17. Hayden, M.S.; Ghosh, S. Shared principles in NF-kappaB signaling. Cell 2008, 132, 344–362. [Google Scholar] [CrossRef]
  18. Lawrence, T. The nuclear factor NF-kappaB pathway in inflammation. Cold Spring Harb. Perspect. Biol. 2009, 1, a001651. [Google Scholar] [CrossRef]
  19. Poma, P. NF-κB and Disease. Int. J. Mol. Sci. 2020, 21, 9181. [Google Scholar] [CrossRef]
  20. DiDonato, J.A.; Mercurio, F.; Karin, M. NF-κB and the link between inflammation and cancer. Immunol. Rev. 2012, 246, 379–400. [Google Scholar] [CrossRef]
  21. Oeckinghaus, A.; Hayden, M.S.; Ghosh, S. Crosstalk in NF-κB signaling pathways. Nat. Immunol. 2011, 12, 695–708. [Google Scholar] [CrossRef] [PubMed]
  22. Paul, A.; Edwards, J.; Pepper, C.; Mackay, S. Inhibitory-κB Kinase (IKK) α and Nuclear Factor-κB (NFκB)-Inducing Kinase (NIK) as Anti-Cancer Drug Targets. Cells. 2018, 7, 176. [Google Scholar] [CrossRef] [PubMed]
  23. Li, Q.; Verma, I.M. NF-kappaB regulation in the immune system. Nat. Rev. Immunol. 2002, 2, 725–734, Erratum in Nat. Rev. Immunol. 2002, 2, 975. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, Q.; Lenardo, M.J.; Baltimore, D. 30 Years of NF-κB: A Blossoming of Relevance to Human Pathobiology. Cell 2017, 168, 37–57. [Google Scholar] [CrossRef] [PubMed]
  25. Mirzaei, S.; Saghari, S.; Bassiri, F.; Raesi, R.; Zarrabi, A.; Hushmandi, K.; Sethi, G.; Tergaonkar, V. NF-κB as a regulator of cancer metastasis and therapy response: A focus on epithelial-mesenchymal transition. J. Cell. Physiol. 2022, 237, 2770–2795. [Google Scholar] [CrossRef] [PubMed]
  26. Hayden, M.S.; West, A.P.; Ghosh, S. SnapShot: NF-kappaB signaling pathways. Cell 2006, 127, 1286–1287. [Google Scholar] [CrossRef] [PubMed]
  27. Wibisana, J.N.; Okada, M. Encoding and decoding NF-κB nuclear dynamics. Curr. Opin. Cell Biol. 2022, 77, 102103. [Google Scholar] [CrossRef]
  28. Gilmore, T.D.; Wolenski, F.S. NF-κB: Where did it come from and why? Immunol. Rev. 2012, 246, 14–35. [Google Scholar] [CrossRef]
  29. Williams, L.M.; Gilmore, T.D. Looking Down on NF-κB. Mol. Cell Biol. 2020, 40, e00104-20. [Google Scholar] [CrossRef]
  30. Pujari, R.; Hunte, R.; Khan, W.N.; Shembade, N. A20-mediated negative regulation of canonical NF-κB signaling pathway. Immunol. Res. 2013, 57, 166–171. [Google Scholar] [CrossRef]
  31. Pham, A.M.; TenOever, B.R. The IKK Kinases: Operators of Antiviral Signaling. Viruses 2010, 2, 55–72. [Google Scholar] [CrossRef] [PubMed]
  32. Zinatizadeh, M.R.; Schock, B.; Chalbatani, G.M.; Zarandi, P.K.; Jalali, S.A.; Miri, S.R. The Nuclear Factor Kappa B (NF-kB) signaling in cancer development and immune diseases. Genes Dis. 2020, 8, 287–297. [Google Scholar] [CrossRef]
  33. Sun, S.C. The non-canonical NF-κB pathway in immunity and inflammation. Nat. Rev. Immunol. 2017, 17, 545–558. [Google Scholar] [CrossRef] [PubMed]
  34. Sun, S.C. Non-canonical NF-κB signaling pathway. Cell Res. 2011, 21, 71–85. [Google Scholar] [CrossRef]
  35. Yu, H.; Lin, L.; Zhang, Z.; Zhang, H.; Hu, H. Targeting NF-κB pathway for the therapy of diseases: Mechanism and clinical study. Signal Transduct. Target. Ther. 2020, 5, 209. [Google Scholar] [CrossRef] [PubMed]
  36. Hou, Y.; Liang, H.; Rao, E.; Zheng, W.; Huang, X.; Deng, L.; Zhang, Y.; Yu, X.; Xu, M.; Mauceri, H.; et al. Non-canonical NF-κB Antagonizes STING Sensor-Mediated DNA Sensing in Radiotherapy. Immunity 2018, 49, 490–503.e4. [Google Scholar] [CrossRef] [PubMed]
  37. Gilmore, T.D. Introduction to NF-kappaB: Players, pathways, perspectives. Oncogene 2006, 25, 6680–6684. [Google Scholar] [CrossRef]
  38. Struzik, J.; Szulc-Dąbrowska, L. Manipulation of Non-canonical NF-κB Signaling by Non-oncogenic Viruses. Arch. Immunol. Ther. Exp. 2019, 67, 41–48. [Google Scholar] [CrossRef]
  39. Meyerovich, K.; Ortis, F.; Cardozo, A.K. The non-canonical NF-κB pathway and its contribution to β-cell failure in diabetes. J. Mol. Endocrinol. 2018, 61, F1–F6. [Google Scholar] [CrossRef]
  40. Lu, X.; Chen, Q.; Liu, H.; Zhang, X. Interplay Between Non-Canonical NF-κB Signaling and Hepatitis B Virus Infection. Front. Immunol. 2021, 29, 730684. [Google Scholar] [CrossRef]
  41. Noort, A.R.; Tak, P.P.; Tas, S.W. Non-canonical NF-κB signaling in rheumatoid arthritis: Dr Jekyll and Mr Hyde? Arthritis Res. Ther. 2015, 17, 15. [Google Scholar] [CrossRef] [PubMed]
  42. Morgan, D.; Garg, M.; Tergaonkar, V.; Tan, S.Y.; Sethi, G. Pharmacological significance of the non-canonical NF-κB pathway in tumorigenesis. Biochim. Biophys. Acta Rev. Cancer 2020, 1874, 188449. [Google Scholar] [CrossRef] [PubMed]
  43. Trares, K.; Ackermann, J.; Koch, I. The canonical and non-canonical NF-κB pathways and their crosstalk: A comparative study based on Petri nets. Biosystems 2022, 211, 104564. [Google Scholar] [CrossRef] [PubMed]
  44. Tao, L.; Ren, X.; Zhai, W.; Chen, Z. Progress and Prospects of Non-Canonical NF-κB Signaling Pathway in the Regulation of Liver Diseases. Molecules 2022, 27, 4275. [Google Scholar] [CrossRef] [PubMed]
  45. O’Donnell, A.; Pepper, C.; Mitchell, S.; Pepper, A. NF-kB and the CLL microenvironment. Front. Oncol. 2023, 30, 1169397. [Google Scholar] [CrossRef] [PubMed]
  46. Shen, R.R.; Hahn, W.C. Emerging roles for the non-canonical IKKs in cancer. Oncogene 2011, 30, 631–641. [Google Scholar] [CrossRef] [PubMed]
  47. Pflug, K.M.; Sitcheran, R. Targeting NF-κB-Inducing Kinase (NIK) in Immunity, Inflammation, and Cancer. Int. J. Mol. Sci. 2020, 21, 8470. [Google Scholar] [CrossRef]
  48. Gu, M.; Zhou, X.; Sohn, J.H.; Zhu, L.; Jie, Z.; Yang, J.Y.; Zheng, X.; Xie, X.; Yang, J.; Shi, Y.; et al. NF-κB-inducing kinase maintains T cell metabolic fitness in antitumor immunity. Nat. Immunol. 2021, 22, 193–204, Erratum in Nat Immunol. 2021 Feb 11. [Google Scholar] [CrossRef]
  49. Xiao, P.; Takiishi, T.; Violato, N.M.; Licata, G.; Dotta, F.; Sebastiani, G.; Marselli, L.; Singh, S.P.; Sze, M.; Van Loo, G.; et al. NF-κB-inducing kinase (NIK) is activated in pancreatic β-cells but does not contribute to the development of diabetes. Cell Death Dis. 2022, 13, 476. [Google Scholar] [CrossRef]
  50. Cheng, J.; Feng, X.; Li, Z.; Zhou, F.; Yang, J.M.; Zhao, Y. Pharmacological inhibition of NF-κB-inducing kinase (NIK) with small molecules for the treatment of human diseases. RSC Med. Chem. 2021, 12, 552–565. [Google Scholar] [CrossRef]
  51. Wang, B.; Shen, J. NF-κB Inducing Kinase Regulates Intestinal Immunity and Homeostasis. Front. Immunol. 2022, 27, 895636. [Google Scholar] [CrossRef] [PubMed]
  52. Zhang, Z.; Zhong, X.; Shen, H.; Sheng, L.; Liangpunsakul, S.; Lok, A.S.; Omary, M.B.; Wang, S.; Rui, L. Biliary NIK promotes ductular reaction and liver injury and fibrosis in mice. Nat. Commun. 2022, 13, 5111. [Google Scholar] [CrossRef] [PubMed]
  53. Choudhary, S.; Sinha, S.; Zhao, Y.; Banerjee, S.; Sathyanarayana, P.; Shahani, S.; Sherman, V.; Tilton, R.G.; Bajaj, M. NF-kappaB-inducing kinase (NIK) mediates skeletal muscle insulin resistance: Blockade by adiponectin. Endocrinology 2011, 152, 3622–3627. [Google Scholar] [CrossRef] [PubMed]
  54. Thu, Y.M.; Su, Y.; Yang, J.; Splittgerber, R.; Na, S.; Boyd, A.; Mosse, C.; Simons, C.; Richmond, A. NF-κB inducing kinase (NIK) modulates melanoma tumorigenesis by regulating expression of pro-survival factors through the β-catenin pathway. Oncogene 2012, 31, 2580–2592. [Google Scholar] [CrossRef] [PubMed]
  55. Fry, C.S.; Nayeem, S.Z.; Dillon, E.L.; Sarkar, P.S.; Tumurbaatar, B.; Urban, R.J.; Wright, T.J.; Sheffield-Moore, M.; Tilton, R.G.; Choudhary, S. Glucocorticoids increase skeletal muscle NF-κB inducing kinase (NIK): Links to muscle atrophy. Physiol. Rep. 2016, 4, e13014. [Google Scholar] [CrossRef]
  56. Hahn, M.; Macht, A.; Waisman, A.; Hövelmeyer, N. NF-κB-inducing kinase is essential for B-cell maintenance in mice. Eur. J. Immunol. 2016, 46, 732–741. [Google Scholar] [CrossRef]
  57. Chung, S.; Sundar, I.K.; Hwang, J.W.; Yull, F.E.; Blackwell, T.S.; Kinnula, V.L.; Bulger, M.; Yao, H.; Rahman, I. NF-κB inducing kinase, NIK mediates cigarette smoke/TNFα-induced histone acetylation and inflammation through differential activation of IKKs. PLoS ONE 2011, 6, e23488. [Google Scholar] [CrossRef]
  58. Yamamoto, M.; Ito, T.; Shimizu, T.; Ishida, T.; Semba, K.; Watanabe, S.; Yamaguchi, N.; Inoue, J. 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]
  59. Davis, J.L.; Thaler, R.; Cox, L.; Ricci, B.; Zannit, H.M.; Wan, F.; Faccio, R.; Dudakovic, A.; van Wijnen, A.J.; Veis, D.J. Constitutive activation of NF-κB inducing kinase (NIK) in the mesenchymal lineage using Osterix (Sp7)- or Fibroblast-specific protein 1 (S100a4)-Cre drives spontaneous soft tissue sarcoma. PLoS ONE 2021, 16, e0254426. [Google Scholar] [CrossRef]
  60. Xiong, Y.; Torsoni, A.S.; Wu, F.; Shen, H.; Liu, Y.; Zhong, X.; Canet, M.J.; Shah, Y.M.; Omary, M.B.; Liu, Y.; et al. Hepatic NF-kB-inducing kinase (NIK) suppresses mouse liver regeneration in acute and chronic liver diseases. Elife 2018, 2, e34152. [Google Scholar] [CrossRef]
  61. Shinzawa, M.; Konno, H.; Qin, J.; Akiyama, N.; Miyauchi, M.; Ohashi, H.; Miyamoto-Sato, E.; Yanagawa, H.; Akiyama, T.; Inoue, J. Catalytic subunits of the phosphatase calcineurin interact with NF-κB-inducing kinase (NIK) and attenuate NIK-dependent gene expression. Sci. Rep. 2015, 1, 10758. [Google Scholar] [CrossRef] [PubMed]
  62. Vazquez-Santillan, K.; Melendez-Zajgla, J.; Jimenez-Hernandez, L.E.; Gaytan-Cervantes, J.; Muñoz-Galindo, L.; Piña-Sanchez, P.; Martinez-Ruiz, G.; Torres, J.; Garcia-Lopez, P.; Gonzalez-Torres, C.; et al. NF-kappaΒ-inducing kinase regulates stem cell phenotype in breast cancer. Sci. Rep. 2016, 23, 37340, Erratum in Sci. Rep. 2017, 23, 44971. [Google Scholar] [CrossRef] [PubMed]
  63. Jiang, B.; Shen, H.; Chen, Z.; Yin, L.; Zan, L.; Rui, L. Carboxyl terminus of HSC70-interacting protein (CHIP) down-regulates NF-κB-inducing kinase (NIK) and suppresses NIK-induced liver injury. J. Biol. Chem. 2015, 290, 11704–11714. [Google Scholar] [CrossRef] [PubMed]
  64. Pflug, K.M.; Lee, D.W.; Keeney, J.N.; Sitcheran, R. NF-κB-inducing kinase maintains mitochondrial efficiency and systemic metabolic homeostasis. Biochim. Biophys. Acta Mol. Basis Dis. 2023, 1869, 166682. [Google Scholar] [CrossRef] [PubMed]
  65. Shen, H.; Ji, Y.; Xiong, Y.; Kim, H.; Zhong, X.; Jin, M.G.; Shah, Y.M.; Omary, M.B.; Liu, Y.; Qi, L.; et al. Medullary thymic epithelial NF-kB-inducing kinase (NIK)/IKKα pathway shapes autoimmunity and liver and lung homeostasis in mice. Proc. Natl. Acad. Sci. USA 2019, 116, 19090–19097. [Google Scholar] [CrossRef] [PubMed]
  66. Thu, Y.M.; Richmond, A. NF-κB inducing kinase: A key regulator in the immune system and in cancer. Cytokine Growth Factor. Rev. 2010, 21, 213–226. [Google Scholar] [CrossRef] [PubMed]
  67. Shen, H.; Sheng, L.; Chen, Z.; Jiang, L.; Su, H.; Yin, L.; Omary, M.B.; Rui, L. Mouse hepatocyte overexpression of NF-κB-inducing kinase (NIK) triggers fatal macrophage-dependent liver injury and fibrosis. Hepatology 2014, 60, 2065–2076. [Google Scholar] [CrossRef]
  68. Sheng, L.; Zhou, Y.; Chen, Z.; Ren, D.; Cho, K.W.; Jiang, L.; Shen, H.; Sasaki, Y.; Rui, L. NF-κB–inducing kinase (NIK) promotes hyperglycemia and glucose intolerance in obesity by augmenting glucagon action. Nat. Med. 2012, 18, 943–949. [Google Scholar] [CrossRef]
  69. Hoberg, J.E.; Yeung, F.; Mayo, M.W. SMRT derepression by the IkappaB kinase alpha: A prerequisite to NF-kappaB transcription and survival. Mol. Cell 2004, 16, 245–255. [Google Scholar] [CrossRef]
  70. Hoberg, J.E.; Popko, A.E.; Ramsey, C.S.; Mayo, M.W. IkappaB kinase alpha-mediated derepression of SMRT potentiates acetylation of RelA/p65 by p300. Mol. Cell Biol. 2006, 26, 457–471. [Google Scholar] [CrossRef]
  71. Fernández-Majada, V.; Aguilera, C.; Villanueva, A.; Vilardell, F.; Robert-Moreno, A.; Aytés, A.; Real, F.X.; Capella, G.; Mayo, M.W.; Espinosa, L.; et al. Nuclear IKK activity leads to dysregulated notch-dependent gene expression in colorectal cancer. Proc. Natl. Acad. Sci. USA 2007, 104, 276–281, Erratum in Proc. Natl. Acad. Sci. USA 2008, 105, 5650. [Google Scholar] [CrossRef] [PubMed]
  72. Fernández-Majada, V.; Pujadas, J.; Vilardell, F.; Capella, G.; Mayo, M.W.; Bigas, A.; Espinosa, L. Aberrant cytoplasmic localization of N-CoR in colorectal tumors. Cell Cycle 2007, 6, 1748–1752. [Google Scholar] [CrossRef] [PubMed]
  73. Margalef, P.; Fernández-Majada, V.; Villanueva, A.; Garcia-Carbonell, R.; Iglesias, M.; López, L.; Martínez-Iniesta, M.; Villà-Freixa, J.; Mulero, M.C.; Andreu, M.; et al. A truncated form of IKKα is responsible for specific nuclear IKK activity in colorectal cancer. Cell Rep. 2012, 2, 840–854. [Google Scholar] [CrossRef] [PubMed]
  74. Park, E.; Liu, B.; Xia, X.; Zhu, F.; Jami, W.B.; Hu, Y. Role of IKKα in skin squamous cell carcinomas. Future Oncol. 2011, 7, 123–134. [Google Scholar] [CrossRef] [PubMed]
  75. Xie, Y.; Xie, K.; Gou, Q.; Chen, N. IκB kinase α functions as a tumor suppressor in epithelial-derived tumors through an NF-κB-independent pathway (Review). Oncol. Rep. 2015, 34, 2225–2232. [Google Scholar] [CrossRef] [PubMed]
  76. Liu, S.; Chen, Z.; Zhu, F.; Hu, Y. IκB kinase alpha and cancer. J. Interferon Cytokine Res. 2012, 32, 152–158. [Google Scholar] [CrossRef]
  77. Zhu, F.; Park, E.; Liu, B.; Xia, X.; Fischer, S.M.; Hu, Y. Critical role of IkappaB kinase alpha in embryonic skin development and skin carcinogenesis. Histol. Histopathol. 2009, 24, 265–271. [Google Scholar] [CrossRef]
  78. Tanese, K. Diagnosis and Management of Basal Cell Carcinoma. Curr. Treat. Options Oncol. 2019, 20, 13. [Google Scholar] [CrossRef]
  79. Molin, S.C.; Grgic, M.; Ruzicka, T.; Herzinger, T. Silencing of the cell cycle checkpoint gene 14-3-3σ in basal cell carcinomas correlates with reduced expression of IKK-α. J. Eur. Acad. Dermatol. Venereol. 2014, 28, 1113–1116. [Google Scholar] [CrossRef]
  80. Mahato, R.; Qin, B.; Cheng, K. Blocking IKKα expression inhibits prostate cancer invasiveness. Pharm. Res. 2011, 28, 1357–1369. [Google Scholar] [CrossRef]
  81. Montes, M.; MacKenzie, L.; McAllister, M.J.; Roseweir, A.; McCall, P.; Hatziieremia, S.; Underwood, M.A.; Boyd, M.; Paul, A.; Plevin, R.; et al. Determining the prognostic significance of IKKα in prostate cancer. Prostate 2020, 80, 1188–1202. [Google Scholar] [CrossRef] [PubMed]
  82. Nguyen, D.P.; Li, J.; Yadav, S.S.; Tewari, A.K. Recent insights into NF-κB signalling pathways and the link between inflammation and prostate cancer. BJU Int. 2014, 114, 168–176. [Google Scholar] [CrossRef] [PubMed]
  83. Man, J.; Zhang, X. CUEDC2: An emerging key player in inflammation and tumorigenesis. Protein Cell 2011, 2, 699–703. [Google Scholar] [CrossRef] [PubMed]
  84. Sau, A.; Lau, R.; Cabrita, M.A.; Nolan, E.; Crooks, P.A.; Visvader, J.E.; Pratt, M.A. Persistent Activation of NF-κB in BRCA1-Deficient Mammary Progenitors Drives Aberrant Proliferation and Accumulation of DNA Damage. Cell Stem Cell 2016, 19, 52–65. [Google Scholar] [CrossRef] [PubMed]
  85. Liao, J.; Qin, Q.H.; Lv, F.Y.; Huang, Z.; Lian, B.; Wei, C.Y.; Mo, Q.G.; Tan, Q.X. IKKα inhibition re-sensitizes acquired adriamycin-resistant triple negative breast cancer cells to chemotherapy-induced apoptosis. Sci. Rep. 2023, 13, 6211. [Google Scholar] [CrossRef] [PubMed]
  86. Ruocco, M.G.; Karin, M. IKK{beta} as a target for treatment of inflammation induced bone loss. Ann. Rheum. Dis. 2005, 64 (Suppl. S4), iv81–iv85. [Google Scholar] [CrossRef]
  87. Page, A.; Navarro, M.; Suárez-Cabrera, C.; Bravo, A.; Ramirez, A. Context-Dependent Role of IKKβ in Cancer. Genes 2017, 8, 376. [Google Scholar] [CrossRef]
  88. Gan, J.; Guo, L.; Zhang, X.; Yu, Q.; Yang, Q.; Zhang, Y.; Zeng, W.; Jiang, X.; Guo, M. Anti-inflammatory therapy of atherosclerosis: Focusing on IKKβ. J. Inflamm. 2023, 20, 8. [Google Scholar] [CrossRef]
  89. Zhang, G.; Li, J.; Purkayastha, S.; Tang, Y.; Zhang, H.; Yin, Y.; Li, B.; Liu, G.; Cai, D. Hypothalamic programming of systemic ageing involving IKK-β, NF-κB and GnRH. Nature 2013, 497, 211–216. [Google Scholar] [CrossRef]
  90. Zhou, P.; Hua, F.; Wang, X.; Huang, J.L. Therapeutic potential of IKK-β inhibitors from natural phenolics for inflammation in cardiovascular diseases. Inflammopharmacology 2020, 28, 19–37. [Google Scholar] [CrossRef]
  91. Heyninck, K.; Lahtela-Kakkonen, M.; Van der Veken, P.; Haegeman, G.; Vanden Berghe, W. Withaferin a inhibits NK-kappaB activation by targeting cysteine 179 in IKKbeta. Biochem. Pharmacol. 2014, 91, 501–509. [Google Scholar] [CrossRef]
  92. Vanden Berghe, W.; Sabbe, L.; Kaileh, M.; Haegeman, G.; Heyninck, K. Molecular insight in the multifunctional activities of withaferin a. Biochem. Pharmacol. 2012, 84, 1282–1291. [Google Scholar] [CrossRef] [PubMed]
  93. Kim, B.H.; Roh, E.; Lee, H.Y.; Lee, I.J.; Ahn, B.; Jung, S.H.; Lee, H.; Han, S.B.; Kim, Y. Benzoxathiole derivative blocks lipopolysaccharide-induced nuclear factor-kappaB activation and nuclear fac-tor-kappaB-regulated gene transcription through inactivating inhibitory kappaB kinase beta. Mol. Pharmacol. 2008, 73, 1309–1318. [Google Scholar] [CrossRef] [PubMed]
  94. Dong, T.; Li, C.; Wang, X.; Dian, L.; Zhang, X.; Li, L.; Chen, S.; Cao, R.; Li, L.; Huang, N.; et al. Ainsliadimer a selectively inhibits IKKα/β by covalently binding a conserved cysteine. Nat. Commun. 2015, 6, 6522. [Google Scholar] [CrossRef] [PubMed]
  95. Prescott, J.A.; Cook, S.J. Targeting IKKβ in Cancer: Challenges and Opportunities for the Therapeutic Utilisation of IKKβ Inhibitors. Cells 2018, 7, 115. [Google Scholar] [CrossRef] [PubMed]
  96. Chen, G.; Liu, S.; Pan, R.; Li, G.; Tang, H.; Jiang, M.; Xing, Y.; Jin, F.; Lin, L.; Dong, J. Curcumin Attenuates gp120-Induced Microglial Inflammation by Inhibiting Autophagy via the PI3K Pathway. Cell. Mol. Neurobiol. 2018, 38, 1465–1477. [Google Scholar] [CrossRef] [PubMed]
  97. Yang, H.; Wang, Y.; Jin, S.; Pang, Q.; Shan, A.; Feng, X. Dietary resveratrol alleviated lipopolysaccharide-induced ileitis through Nrf2 and NF-κB signalling pathways in ducks (Anas platyrhynchos). J. Anim. Physiol. Anim. Nutr. 2022, 106, 1306–1320. [Google Scholar] [CrossRef]
  98. Li, A.; Lin, C.; Xie, F.; Jin, M.; Lin, F. Berberine Ameliorates Insulin Resistance by Inhibiting IKK/NF-κB, JNK, and IRS-1/AKT Signaling Pathway in Liver of Gestational Diabetes Mellitus Rats. Metab. Syndr. Relat. Disord. 2022, 20, 480–488. [Google Scholar] [CrossRef]
  99. Wheeler, D.S.; Catravas, J.D.; Odoms, K.; Denenberg, A.; Malhotra, V.; Wong, H.R. Epigallocatechin-3-gallate, a green tea-derived polyphenol, inhibits IL-1 beta-dependent proinflammatory signal transduction in cultured respiratory epithelial cells. J. Nutr. 2004, 134, 1039–1044. [Google Scholar] [CrossRef]
  100. Veerappan, K.; Natarajan, S.; Ethiraj, P.; Vetrivel, U.; Samuel, S. Inhibition of IKKβ by celastrol and its analogues—An in silico and in vitro approach. Pharm. Biol. 2017, 55, 368–373. [Google Scholar] [CrossRef]
  101. Rauert-Wunderlich, H.; Siegmund, D.; Maier, E.; Giner, T.; Bargou, R.C.; Wajant, H.; Stühmer, T. The IKK inhibitor Bay 11-7082 induces cell death independent from inhibition of activation of NFκB transcription factors. PLoS ONE 2013, 8, e59292. [Google Scholar] [CrossRef] [PubMed]
  102. Lung, H.L.; Kan, R.; Chau, W.Y.; Man, O.Y.; Mak, N.K.; Fong, C.H.; Shuen, W.H.; Tsao, S.W.; Lung, M.L. The anti-tumor function of the IKK inhibitor PS1145 and high levels of p65 and KLF4 are associated with the drug resistance in nasopharyngeal carcinoma cells. Sci. Rep. 2019, 9, 12064. [Google Scholar] [CrossRef] [PubMed]
  103. Sachse, F.; Becker, K.; Basel, T.J.; Weiss, D.; Rudack, C. IKK-2 inhibitor TPCA-1 represses nasal epithelial inflammation in vitro. Rhinology 2011, 49, 168–173. [Google Scholar] [CrossRef] [PubMed]
  104. Waga, K.; Yamaguchi, M.; Miura, S.; Nishida, T.; Itai, A.; Nakanishi, R.; Kashiwakura, I. IKKβ Inhibitor IMD-0354 Attenuates Radiation Damage in Whole-body X-Irradiated Mice. Oxid. Med. Cell. Longev. 2019, 27, 5340290. [Google Scholar] [CrossRef] [PubMed]
  105. Clark, K.; Peggie, M.; Plater, L.; Sorcek, R.J.; Young, E.R.; Madwed, J.B.; Hough, J.; McIver, E.G.; Cohen, P. Novel cross-talk within the IKK family controls innate immunity. Biochem. J. 2011, 434, 93–104. [Google Scholar] [CrossRef] [PubMed]
  106. Li, Q.T.; Van Antwerp, D.; Mercurio, F.; Lee, K.F.; Verma, I.M. Severe liver degeneration in mice lacking the IkappaB kinase 2 gene. Science 1999, 284, 321–325. [Google Scholar] [CrossRef] [PubMed]
  107. Llona-Minguez, S.; Baiget, J.; Mackay, S.P. Small-molecule inhibitors of IκB kinase (IKK) and IKK-related kinases. Pharm. Pat. Anal. 2013, 2, 481–498. [Google Scholar] [CrossRef]
  108. Asamitsu, K.; Yamaguchi, T.; Nakata, K.; Hibi, Y.; Victoriano, A.F.B.; Imai, K.; Onozaki, K.; Kitade, Y.; Okamoto, T. Inhibition of human immunodeficiency virus type 1 replication by blocking IkappaB Kinase with noraristeromycin. J. Biochem. 2008, 144, 581–589. [Google Scholar] [CrossRef]
  109. Bhargava, P.; Malik, V.; Liu, Y.; Ryu, J.; Kaul, S.C.; Sundar, D.; Wadhwa, R. Molecular Insights Into Withaferin-A-Induced Senescence: Bioinformatics and Experimental Evidence to the Role of NFκB and CARF. J. Gerontol. A Biol. Sci. Med. Sci. 2019, 74, 183–191. [Google Scholar] [CrossRef]
  110. Duan, X.; Ponomareva, L.; Veeranki, S.; Panchanathan, R.; Dickerson, E.; Choubey, D. Differential roles for the interferon-inducible IFI16 and AIM2 innate immune sensors for cytosolic DNA in cellular senescence of human fibroblasts. Mol. Cancer Res. 2011, 9, 589–602. [Google Scholar] [CrossRef]
  111. Fafián-Labora, J.A.; O’Loghlen, A. NF-κB/IKK activation by small extracellular vesicles within the SASP. Aging Cell 2021, 20, e13426. [Google Scholar] [CrossRef] [PubMed]
  112. Li, T.; Yan, B.; Xiao, X.; Zhou, L.; Zhang, J.; Yuan, Q.; Shan, L.; Wu, H.; Efferth, T. Onset of p53/NF-κB signaling crosstalk in human melanoma cells in response to anti-cancer theabrownin. FASEB J. 2022, 36, e22426. [Google Scholar] [CrossRef] [PubMed]
  113. Choubey, D.; Panchanathan, R. IFI16, an amplifier of DNA-damage response: Role in cellular senescence and aging-associated inflammatory diseases. Ageing Res. Rev. 2016, 28, 27–36. [Google Scholar] [CrossRef] [PubMed]
  114. Lorenz, J.; Zahlten, J.; Pollok, I.; Lippmann, J.; Scharf, S.; N’Guessan, P.D.; Opitz, B.; Flieger, A.; Suttorp, N.; Hippenstiel, S.; et al. Legionella pneumophila-induced IκBζ-dependent expression of interleukin-6 in lung epithelium. Eur. Respir. J. 2011, 37, 648–657. [Google Scholar] [CrossRef] [PubMed]
  115. Cuollo, L.; Antonangeli, F.; Santoni, A.; Soriani, A. The Senescence-Associated Secretory Phenotype (SASP) in the Challenging Future of Cancer Therapy and Age-Related Diseases. Biology 2020, 9, 485. [Google Scholar] [CrossRef] [PubMed]
  116. Burton, D.G.; Faragher, R.G. Cellular senescence: From growth arrest to immunogenic conversion. Age. 2015, 37, 27. [Google Scholar] [CrossRef] [PubMed]
  117. Moiseeva, O.; Deschenes-Simard, X.; St-Germain, E.; Igelmann, S.; Huot, G.; Cadar, A.E.; Bourdeau, V.; Pollak, M.N.; Ferbeyre, G. Metformin inhibits the senescence-associated secretory phenotype by interfering with IKK/NF-kappaB activation. Aging Cell 2013, 12, 489–498. [Google Scholar] [CrossRef]
  118. Birch, J.; Gil, J. Senescence and the SASP: Many therapeutic avenues. Genes Dev. 2020, 34, 1565–1576, PMID: 33262144; PMCID: PMC7706700. [Google Scholar] [CrossRef]
  119. Tilstra, J.S.; Robinson, A.R.; Wang, J.; Gregg, S.Q.; Clauson, C.L.; Reay, D.P.; Nasto, L.A.; St Croix, C.M.; Usas, A.; Vo, N.; et al. NF-kappaB inhibition delays DNA damage-induced senescence and aging in mice. J. Clin. Investig. 2012, 122, 2601–2612. [Google Scholar] [CrossRef]
  120. Song, W.; Li, D.; Tao, L.; Luo, Q.; Chen, L. Solute carrier transporters: The metabolic gatekeepers of immune cells. Acta Pharm. Sin. B 2020, 10, 61–78. [Google Scholar] [CrossRef]
  121. Heuser, C.; Gotot, J.; Piotrowski, E.C.; Philipp, M.S.; Courrèges, C.J.F.; Otte, M.S.; Guo, L.; Schmid-Burgk, J.L.; Hornung, V.; Heine, A.; et al. Prolonged IKKβ Inhibition Improves Ongoing CTL Antitumor Responses by Incapacitating Regulatory T Cells. Cell Rep. 2017, 21, 578–586. [Google Scholar] [CrossRef] [PubMed]
  122. Senftleben, U.; Li, Z.W.; Baud, V.; Karin, M. IKKbeta is essential for protecting T cells from TNFalpha-induced apoptosis. Immunity 2001, 14, 217–230. [Google Scholar] [CrossRef] [PubMed]
  123. Deng, H.; Mao, G.; Zhang, J.; Wang, Z.; Li, D. IKK antagonizes activation-induced cell death of CD4+ T cells in aged mice via inhibition of JNK activation. Mol. Immunol. 2010, 48, 287–293. [Google Scholar] [CrossRef] [PubMed]
  124. Blonska, M.; Joo, D.; Nurieva, R.I.; Zhao, X.; Chiao, P.; Sun, S.C.; Dong, C.; Lin, X. Activation of the transcription factor c-Maf in T cells is dependent on the CARMA1-IKKβ signaling cascade. Sci. Signal. 2013, 6, ra110. [Google Scholar] [CrossRef] [PubMed]
  125. Krishna, S.; Xie, D.; Gorentla, B.; Shin, J.; Gao, J.; Zhong, X.P. Chronic activation of the kinase IKKβ impairs T cell function and survival. J. Immunol. 2012, 189, 1209–1219. [Google Scholar] [CrossRef] [PubMed]
  126. Miller, M.L.; Mashayekhi, M.; Chen, L.; Zhou, P.; Liu, X.; Michelotti, M.; Tramontini Gunn, N.; Powers, S.; Zhu, X.; Evaristo, C.; et al. Basal NF-κB controls IL-7 responsiveness of quiescent naïve T cells. Proc. Natl. Acad. Sci. USA 2014, 111, 7397–7402. [Google Scholar] [CrossRef] [PubMed]
  127. Jeucken, K.C.M.; van Rooijen, C.C.N.; Kan, Y.Y.; Kocken, L.A.; Jongejan, A.; van Steen, A.C.I.; van Buul, J.D.; Olsson, H.K.; van Hamburg, J.P.; Tas, S.W. Differential Contribution of NF-κB Signaling Pathways to CD4+ Memory T Cell Induced Activation of Endothelial Cells. Front. Immunol. 2022, 13, 860327. [Google Scholar] [CrossRef]
  128. Chuang, H.C.; Tsai, C.Y.; Hsueh, C.H.; Tan, T.H. GLK-IKKβ signaling induces dimerization and translocation of the AhR-RORγt complex in IL-17A induction and autoimmune disease. Sci. Adv. 2018, 4, eaat5401. [Google Scholar] [CrossRef]
  129. Evaristo, C.; Spranger, S.; Barnes, S.E.; Miller, M.L.; Molinero, L.L.; Locke, F.L.; Gajewski, T.F.; Alegre, M.L. Cutting Edge: Engineering Active IKKβ in T Cells Drives Tumor Rejection. J. Immunol. 2016, 196, 2933–2938. [Google Scholar] [CrossRef]
  130. Ding, W.; Wagner, J.A.; Granstein, R.D. CGRP, PACAP, and VIP modulate Langerhans cell function by inhibiting NF-kappaB activation. J. Investig. Dermatol. 2007, 127, 2357–2367. [Google Scholar] [CrossRef]
  131. Bosch, N.C.; Voll, R.E.; Voskens, C.J.; Gross, S.; Seliger, B.; Schuler, G.; Schaft, N.; Dörrie, J. NF-κB activation triggers NK-cell stimulation by monocyte-derived dendritic cells. Ther. Adv. Med. Oncol. 2019, 11, 1758835919891622. [Google Scholar] [CrossRef] [PubMed]
  132. Yang, J.; Hawkins, O.E.; Barham, W.; Gilchuk, P.; Boothby, M.; Ayers, G.D.; Joyce, S.; Karin, M.; Yull, F.E.; Richmond, A. Myeloid IKKβ promotes antitumor immunity by modulating CCL11 and the innate immune response. Cancer Res. 2014, 74, 7274–7284. [Google Scholar] [CrossRef] [PubMed]
  133. Lopez-Pelaez, M.; Lamont, D.J.; Peggie, M.; Shpiro, N.; Gray, N.S.; Cohen, P. Protein kinase IKKβ-catalyzed phosphorylation of IRF5 at Ser462 induces its dimerization and nuclear translocation in myeloid cells. Proc. Natl. Acad. Sci. USA 2014, 111, 17432–17437. [Google Scholar] [CrossRef] [PubMed]
  134. Baratin, M.; Foray, C.; Demaria, O.; Habbeddine, M.; Pollet, E.; Maurizio, J.; Verthuy, C.; Davanture, S.; Azukizawa, H.; Flores-Langarica, A.; et al. Homeostatic NF-κB Signaling in Steady-State Migratory Dendritic Cells Regulates Immune Homeostasis and Tolerance. Immunity 2015, 42, 627–639. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, X.; Wang, J.; Zheng, H.; Xie, M.; Hopewell, E.L.; Albrecht, R.A.; Nogusa, S.; García-Sastre, A.; Balachandran, S.; Beg, A.A. Differential requirement for the IKKβ/NF-κB signaling module in regulating TLR- versus RLR-induced type 1 IFN expression in dendritic cells. J. Immunol. 2014, 193, 2538–2545. [Google Scholar] [CrossRef]
  136. Chow, K.T.; Wilkins, C.; Narita, M.; Green, R.; Knoll, M.; Loo, Y.M.; Gale, M., Jr. Differential and Overlapping Immune Programs Regulated by IRF3 and IRF5 in Plasmacytoid Dendritic Cells. J. Immunol. 2018, 201, 3036–3050. [Google Scholar] [CrossRef]
  137. Karin, M.; Ben-Neriah, Y. Phosphorylation meets ubiquitination: The control of NFkappa]B activity. Annu. Rev. Immunol. 2000, 18, 621–663. [Google Scholar] [CrossRef]
  138. Erol, A. IKK-mediated CYLD phosphorylation and cellular redox activity. Mol. Med. 2022, 28, 14. [Google Scholar] [CrossRef]
  139. Pradère, J.P.; Hernandez, C.; Koppe, C.; Friedman, R.A.; Luedde, T.; Schwabe, R.F. Negative regulation of NF-κB p65 activity by serine 536 phosphorylation. Sci. Signal. 2016, 9, ra85. [Google Scholar] [CrossRef]
  140. Zhang, J.; Zhao, R.; Yu, C.; Bryant, C.L.N.; Wu, K.; Liu, Z.; Ding, Y.; Zhao, Y.; Xue, B.; Pan, Z.Q.; et al. IKK-Mediated Regulation of the COP9 Signalosome via Phosphorylation of CSN5. J. Proteome Res. 2020, 19, 1119–1130. [Google Scholar] [CrossRef]
  141. Blanchett, S.; Dondelinger, Y.; Barbarulo, A.; Bertrand, M.J.M.; Seddon, B. Phosphorylation of RIPK1 serine 25 mediates IKK dependent control of extrinsic cell death in T cells. Front. Immunol. 2022, 1, 1067164. [Google Scholar] [CrossRef] [PubMed]
  142. Zhang, X.; Yu, H.; Zhao, J.; Li, X.; Li, J.; He, J.; Xia, Z.; Zhao, J. IKKϵ negatively regulates RIG-I via direct phosphorylation. J. Med. Virol. 2016, 88, 712–718. [Google Scholar] [CrossRef] [PubMed]
  143. Zhao, P.; Wong, K.I.; Sun, X.; Reilly, S.M.; Uhm, M.; Liao, Z.; Skorobogatko, Y.; Saltiel, A.R. TBK1 at the Crossroads of Inflammation and Energy Homeostasis in Adipose Tissue. Cell 2018, 172, 731–743.e12. [Google Scholar] [CrossRef] [PubMed]
  144. Remoli, A.L.; Sgarbanti, M.; Perrotti, E.; Acchioni, M.; Orsatti, R.; Acchioni, C.; Battistini, A.; Clarke, R.; Marsili, G. IκB kinase-ε-mediated phosphorylation triggers IRF-1 degradation in breast cancer cells. Neoplasia 2020, 22, 459–469. [Google Scholar] [CrossRef] [PubMed]
  145. Amaya, M.; Keck, F.; Bailey, C.; Narayanan, A. The role of the IKK complex in viral infections. Pathog Dis. 2014, 72, 32–44, Epub 2014 Aug 28. PMID: 25082354; PMCID: PMC7108545. [Google Scholar] [CrossRef] [PubMed]
  146. Wang, V.Y.; Li, Y.; Kim, D.; Zhong, X.; Du, Q.; Ghassemian, M.; Ghosh, G. Bcl3 Phosphorylation by Akt, Erk2, and IKK Is Required for Its Transcriptional Activity. Mol. Cell 2017, 67, 484–497.e5. [Google Scholar] [CrossRef] [PubMed]
  147. Ling, L.; Cao, Z.; Goeddel, D.V. NF-kappaB-inducing kinase activates IKK-alpha by phosphorylation of Ser-176. Proc. Natl. Acad. Sci. USA 1998, 95, 3792–3797. [Google Scholar] [CrossRef]
  148. Antonia, R.J.; Baldwin, A.S. IKK promotes cytokine-induced and cancer-associated AMPK activity and attenuates phenformin-induced cell death in LKB1-deficient cells. Sci. Signal. 2018, 11, eaan5850. [Google Scholar] [CrossRef]
  149. Lee, D.F.; Hung, M.C. Advances in targeting IKK and IKK-related kinases for cancer therapy. Clin. Cancer Res. 2008, 14, 5656–5662. [Google Scholar] [CrossRef]
  150. Schomer-Miller, B.; Higashimoto, T.; Lee, Y.K.; Zandi, E. Regulation of IkappaB kinase (IKK) complex by IKKgamma-dependent phosphorylation of the T-loop and C terminus of IKKbeta. J. Biol. Chem. 2006, 281, 15268–15276. [Google Scholar] [CrossRef]
  151. Hehner, S.P.; Hofmann, T.G.; Ushmorov, A.; Dienz, O.; Wing-Lan Leung, I.; Lassam, N.; Scheidereit, C.; Dröge, W.; Schmitz, M.L. Mixed-lineage kinase 3 delivers CD3/CD28-derived signals into the IkappaB kinase complex. Mol. Cell Biol. 2000, 20, 2556–2568. [Google Scholar] [CrossRef] [PubMed]
  152. Li, N.; Banin, S.; Ouyang, H.; Li, G.C.; Courtois, G.; Shiloh, Y.; Karin, M.; Rotman, G. ATM is required for IkappaB kinase (IKKk) activation in response to DNA double strand breaks. J. Biol. Chem. 2001, 276, 8898–8903. [Google Scholar] [CrossRef] [PubMed]
  153. Menden, H.; Tate, E.; Hogg, N.; Sampath, V. LPS-mediated endothelial activation in pulmonary endothelial cells: Role of Nox2-dependent IKK-β phosphorylation. Am. J. Physiol. Lung Cell. Mol. Physiol. 2013, 304, L445–L455. [Google Scholar] [CrossRef]
  154. Li, T.; Wong, V.K.; Jiang, Z.H.; Jiang, S.P.; Liu, Y.; Wang, T.Y.; Yao, X.J.; Su, X.H.; Yan, F.G.; Liu, J.; et al. Mutation of cysteine 46 in IKK-beta increases inflammatory responses. Oncotarget 2015, 6, 31805–31819. [Google Scholar] [CrossRef] [PubMed]
  155. Nomura, F.; Kawai, T.; Nakanishi, K.; Akira, S. NF-kappaB activation through IKK-i-dependent I-TRAF/TANK phosphorylation. Genes Cells 2000, 5, 191–202. [Google Scholar] [CrossRef] [PubMed]
  156. Motolani, A.; Martin, M.; Sun, M.; Lu, T. Phosphorylation of the Regulators, a Complex Facet of NF-κB Signaling in Cancer. Biomolecules 2020, 11, 15. [Google Scholar] [CrossRef] [PubMed]
  157. Chariot, A. The NF-kappaB-independent functions of IKK subunits in immunity and cancer. Trends Cell Biol. 2009, 19, 404–413. [Google Scholar] [CrossRef] [PubMed]
  158. Whitesid, S.T.; Ernst, M.K.; LeBail, O.; Laurent-Winter, C.; Rice, N.; Israël, A. N- and C-terminal sequences control degradation of MAD3/I kappa B alpha in response to inducers of NF-kappa B activity. Mol. Cell Biol. 1995, 15, 5339–5345. [Google Scholar] [CrossRef]
  159. Palkowitsch, L.; Leidner, J.; Ghosh, S.; Marienfeld, R.B. Phosphorylation of serine 68 in the IkappaB kinase (IKK)-binding domain of NEMO interferes with the structure of the IKK complex and tumor necrosis factor-alpha-induced NF-kappaB activity. J. Biol. Chem. 2008, 283, 76–86. [Google Scholar] [CrossRef]
  160. Zandi, E.; Rothwarf, D.M.; Delhase, M.; Hayakawa, M.; Karin, M. The IkappaB kinase complex (IKK) contains two kinase subunits, IKKalpha and IKKbeta, necessary for IkappaB phosphorylation and NF-kappaB activation. Cell 1997, 91, 243–252. [Google Scholar] [CrossRef] [PubMed]
  161. Garbati, M.R.; Gilmore, T.D. Ser484 and Ser494 in REL are the major sites of IKK phosphorylation in vitro: Evidence that IKK does not directly enhance GAL4-REL transactivation. Gene Expr. 2008, 14, 195–205. [Google Scholar] [CrossRef] [PubMed]
  162. Hinz, M.; Scheidereit, C. The IκB kinase complex in NF-κB regulation and beyond. EMBO Rep. 2014, 15, 46–61. [Google Scholar] [CrossRef] [PubMed]
  163. Chen, Z.J. Ubiquitination in signaling to and activation of IKK. Immunol. Rev. 2012, 246, 95–106. [Google Scholar] [CrossRef] [PubMed]
  164. Napetschnig, J.; Wu, H. Molecular basis of NF-κB signaling. Annu. Rev. Biophys. 2013, 42, 443–468. [Google Scholar] [CrossRef] [PubMed]
  165. Ben, J.; Jiang, B.; Wang, D.; Liu, Q.; Zhang, Y.; Qi, Y.; Tong, X.; Chen, L.; Liu, X.; Zhang, Y.; et al. Major vault protein suppresses obesity and atherosclerosis through inhibiting IKK-NF-κB signaling mediated inflammation. Nat. Commun. 2019, 10, 1801. [Google Scholar] [CrossRef] [PubMed]
  166. Walsh, M.C.; Lee, J.; Choi, Y. Tumor necrosis factor receptor- associated factor 6 (TRAF6) regulation of development, function, and homeostasis of the immune system. Immunol. Rev. 2015, 266, 72–92. [Google Scholar] [CrossRef] [PubMed]
  167. Kanarek, N.; Ben-Neriah, Y. Regulation of NF-κB by ubiquitination and degradation of the IκBs. Immunol. Rev. 2012, 246, 77–94. [Google Scholar] [CrossRef]
  168. Karim, Z.A.; Vemana, H.P.; Khasawneh, F.T. MALT1-ubiquitination triggers non-genomic NF-κB/IKK signaling upon platelet activation. PLoS ONE 2015, 10, e0119363. [Google Scholar] [CrossRef]
  169. Vallabhapurapu, S.; Matsuzawa, A.; Zhang, W. 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]
  170. Ji, Y.X.; Zhang, P.; Zhang, X.J.; Zhao, Y.C.; Deng, K.Q.; Jiang, X.; Wang, P.X.; Huang, Z.; Li, H. The ubiquitin E3 ligase TRAF6 exacerbates pathological cardiac hypertrophy via TAK1-dependent signalling. Nat. Commun. 2016, 1, 11267. [Google Scholar] [CrossRef]
  171. Ohtake, F.; Saeki, Y.; Ishido, S.; Kanno, J.; Tanaka, K. The K48–K63 Branched Ubiquitin Chain Regulates NF-κB Signaling. Mol. Cell 2016, 64, 251–266. [Google Scholar] [CrossRef] [PubMed]
  172. Tang, Y.; Tu, H.; Zhang, J.; Zhao, X.; Wang, Y.; Qin, J.; Lin, X. K63-linked ubiquitination regulates RIPK1 kinase activity to prevent cell death during embryogenesis and inflammation. Nat. Commun. 2019, 10, 4157. [Google Scholar] [CrossRef] [PubMed]
  173. Tarantino, N.; Tinevez, J.Y.; Crowell, E.F.; Boisson, B.; Henriques, R.; Mhlanga, M.; Agou, F.; Israël, A.; Laplantine, E. TNF and IL-1 exhibit distinct ubiquitin requirements for inducing NEMO-IKK supramolecular structures. J. Cell Biol. 2014, 204, 231–245. [Google Scholar] [CrossRef] [PubMed]
  174. Kai, X.; Chellappa, V.; Donado, C.; Reyon, D.; Sekigami, Y.; Ataca, D.; Louissaint, A.; Mattoo, H.; Joung, J.K.; Pillai, S. IκB kinase β (IKBKB) mutations in lymphomas that constitutively activate canonical nuclear factor κB (NFκB) signaling. J. Biol. Chem. 2014, 289, 26960–26972. [Google Scholar] [CrossRef] [PubMed]
  175. Spina, V.; Rossi, D. NF-κB deregulation in splenic marginal zone lymphoma. Semin. Cancer Biol. 2016, 39, 61–67. [Google Scholar] [CrossRef]
  176. Barwick, B.G.; Gupta, V.A.; Vertino, P.M.; Boise, L.H. Cell of Origin and Genetic Alterations in the Pathogenesis of Multiple Myeloma. Front. Immunol. 2019, 21, 1121. [Google Scholar] [CrossRef]
  177. Sasaki, Y.; Calado, D.P.; Derudder, E.; Zhang, B.; Shimizu, Y.; Mackay, F.; Nishikawa, S.; Rajewsky, K.; Schmidt-Supprian, M. NIK overexpression amplifies, whereas ablation of its TRAF3-binding domain replaces BAFF:BAFF-R-mediated survival signals in B cells. Proc. Natl. Acad. Sci. USA 2008, 105, 10883–10888. [Google Scholar] [CrossRef]
  178. Gatla, H.R.; Zou, Y.; Uddin, M.M.; Singha, B.; Bu, P.; Vancura, A.; Vancurova, I. Histone Deacetylase (HDAC) Inhibition Induces IκB Kinase (IKK)-dependent Interleukin-8/CXCL8 Expression in Ovarian Cancer Cells. J. Biol. Chem. 2017, 292, 5043–5054. [Google Scholar] [CrossRef]
  179. Albensi, B.C.; Mattson, M.P. Evidence for the involvement of TNF and NF-kappaB in hippocampal synaptic plasticity. Synapse 2000, 35, 151–159. [Google Scholar] [CrossRef]
  180. Lubin, F.D.; Sweatt, J.D. The IkappaB kinase regulates chromatin structure during reconsolidation of conditioned fear memories. Neuron 2007, 55, 942–957. [Google Scholar] [CrossRef]
  181. Saliminejad, K.; Khorram Khorshid, H.R.; Soleymani Fard, S.; Ghaffari, S.H. An overview of microRNAs: Biology, functions, therapeutics, and analysis methods. J. Cell. Physiol. 2019, 234, 5451–5465. [Google Scholar] [CrossRef] [PubMed]
  182. Yang, Z.; Fang, S.; Di, Y.; Ying, W.; Tan, Y.; Gu, W. Modulation of NF-κB/miR-21/PTEN pathway sensitizes non-small cell lung cancer to cisplatin. PLoS ONE 2015, 10, e0121547. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of the roles of IκB in the canonical and non-canonical NF-κB pathway. The NF-kB pathway involves key proteins such as Tumor Necrosis Factor Receptor-Associated Factor 3 (TRAF3), Precursor protein 100 (p100), V-rel avian reticuloendotheliosis viral oncogene homolog B (RelB), Nuclear Factor NF-kappa-B p52 subunit (p52), Nuclear Factor NF-kappa-B p65 subunit (p65), Cellular Inhibitor of Apoptosis Protein 1/2 (cIAP 1/2), B-cell Activating Factor Receptor (BAFF-R), and Lymphotoxin Beta Receptor (LTβR). Ub indicates ubiquitination, IKK α ((IkappaB kinase alpha). P indicates phosphorylation.
Figure 1. Schematic representation of the roles of IκB in the canonical and non-canonical NF-κB pathway. The NF-kB pathway involves key proteins such as Tumor Necrosis Factor Receptor-Associated Factor 3 (TRAF3), Precursor protein 100 (p100), V-rel avian reticuloendotheliosis viral oncogene homolog B (RelB), Nuclear Factor NF-kappa-B p52 subunit (p52), Nuclear Factor NF-kappa-B p65 subunit (p65), Cellular Inhibitor of Apoptosis Protein 1/2 (cIAP 1/2), B-cell Activating Factor Receptor (BAFF-R), and Lymphotoxin Beta Receptor (LTβR). Ub indicates ubiquitination, IKK α ((IkappaB kinase alpha). P indicates phosphorylation.
Kinasesphosphatases 02 00002 g001
Table 1. Types of IκB kinases.
Table 1. Types of IκB kinases.
Type of IκB KinaseRole in NF-κB PathwayFunction in NF-κB RegulationTargets
IKKα (Inhibitor of κB Kinase Alpha) [7]Non-Canonical NF-κB pathwayPhosphorylates p100, leading to partial proteasomal processing into p52. Initiates non-canonical gene transcription. Involved in cellular senescence.p100
IKKβ (Inhibitor of κB Kinase Beta) [8]Canonical NF-κB pathwayPhosphorylates IκBα and IκBβ, marking them for degradation. Releases p50-RelA dimers for nuclear translocation. Associated with chronic inflammation and tumor promotion.IκBα, IκBβ
IKKε (Inhibitor of κB Kinase Epsilon) [9]Both canonical and non-canonical pathwaysRegulates NF-κB activation, particularly in response to viral infections. Can promote cell survival.-
TBK1 (TANK-binding kinase 1) [10]Non-Canonical NF-κB pathwayActivates IKKα and promotes non-canonical NF-κB signaling. Also involved in antiviral immune responses.IKKα
IKKζ (Inhibitor of κB Kinase Zeta (MAIL)) [11]Both canonical and non-canonical pathwaysModulates NF-κB signaling and immune responses. May play a role in inflammation and autoimmunity.-
NEMO (NF-κB Essential Modulator) [12]Central scaffold proteinActs as an essential scaffold for IKKα and IKKβ, facilitating their activation. Essential for canonical NF-κB activation.IKKα, IKKβ
Table 2. The selected compounds of natural and synthetic origin that were investigated for their effects on IKKβ in various cancer cell lines. They modulate IKKβ activity, leading to a suppression of NF-κB signaling, and the consequent downregulation of genes associated with inflammation, cell survival, and proliferation.
Table 2. The selected compounds of natural and synthetic origin that were investigated for their effects on IKKβ in various cancer cell lines. They modulate IKKβ activity, leading to a suppression of NF-κB signaling, and the consequent downregulation of genes associated with inflammation, cell survival, and proliferation.
Compound NameSource or SynthesisCell Line/OrganismConcentration (μM)Incubation Time (h)Observed Effect on IKKβ/TargetStructure
Curcumin [96]Turmeric (Plant)Various cancer cell lines (e.g., MCF-7, A549)10–5012–48Inhibition of IKKβ phosphorylation and NF-κB activation, leading to reduced pro-inflammatory and pro-survival gene expression.Kinasesphosphatases 02 00002 i001
Resveratrol [97]Red grapes (plant)Human prostate cancer cells (e.g., PC-3)50–10024–72Suppression of IKKβ activity, resulting in reduced NF-κB-mediated transcription and anti-proliferative effects.Kinasesphosphatases 02 00002 i002
Berberine [98]Berberis pantVarious cancer cell lines (e.g., HCT-116, MDA-MB-231)10–10012–48Inhibition of IKKβ phosphorylation, blocking NF-κB activation, and reducing the expression of pro-inflammatory and anti-apoptotic genes.Kinasesphosphatases 02 00002 i003
EGCG (epigallocatechin-3-gallate) [99]Green tea (plant)Various cancer cell lines (e.g., A549, HCT-116)20–10024–48Suppression of IKKβ phosphorylation, leading to decreased NF-κB activity and inhibition of pro-survival and pro-inflammatory pathways.Kinasesphosphatases 02 00002 i004
Celastrol [100]Thunder of god vine (plant)Human beast cancer cells (e.g., MDA-MB-231)0.5–16–24Inhibition of IKKβ activity, blocking NF-κB signaling, and promoting apoptosis in cancer cells.Kinasesphosphatases 02 00002 i005
BAY 11-7082 [101]Synthetic compoundMultiple cancer cell lines (e.g., HeLa, U87)5–202–24Direct inhibition of IKKβ activity, leading to the suppression of NF-κB signaling and the downregulation of pro-survival and pro-inflammatory genes.Kinasesphosphatases 02 00002 i006
PS1145 [102]Synthetic compoundVarious cancer cell lines (e.g., A549, MDA-MB-231)1–104–24Selective inhibition of IKKβ, resulting in the attenuation of NF-κB signaling and the reduction in pro-inflammatory and anti-apoptotic gene expression.Kinasesphosphatases 02 00002 i007
TPCA-1 [103]Synthetic compoundHuman lung cancer cells (e.g., H1299)1–56–24Inhibition of IKKβ kinase activity, leading to the suppression of NF-κB-mediated transcription and anti-proliferative effects.Kinasesphosphatases 02 00002 i008
IMD-0354 [104]Synthetic compoundProstate cancer cells (e.g., PC-3)10–506–48Inhibition of IKKβ activity, resulting in reduced NF-κB signaling and the downregulation of genes associated with cell survival and inflammation.Kinasesphosphatases 02 00002 i009
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

Abdrabou, A.M. The Yin and Yang of IκB Kinases in Cancer. Kinases Phosphatases 2024, 2, 9-27. https://doi.org/10.3390/kinasesphosphatases2010002

AMA Style

Abdrabou AM. The Yin and Yang of IκB Kinases in Cancer. Kinases and Phosphatases. 2024; 2(1):9-27. https://doi.org/10.3390/kinasesphosphatases2010002

Chicago/Turabian Style

Abdrabou, Abdalla M. 2024. "The Yin and Yang of IκB Kinases in Cancer" Kinases and Phosphatases 2, no. 1: 9-27. https://doi.org/10.3390/kinasesphosphatases2010002

Article Metrics

Back to TopTop