Next Article in Journal
Identification of Viruses Infecting Phalaenopsis Orchids Using Nanopore Sequencing and Development of an RT-RPA-CRISPR/Cas12a for Rapid Visual Detection of Nerine Latent Virus
Previous Article in Journal
Application of Aquaporins as Markers in Forensic Pathology: A Systematic Review of the Literature
Previous Article in Special Issue
Circulating isomiRs May Be Superior Biomarkers Compared to Their Corresponding miRNAs: A Pilot Biomarker Study of Using isomiR-Ome to Detect Coronary Calcium-Based Cardiovascular Risk in Patients with NAFLD
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Roles of lncRNAs in NF-κB-Mediated Macrophage Inflammation and Their Implications in the Pathogenesis of Human Diseases

by
Jae-Joon Shin
1,
Jeongkwang Park
1,
Hyeung-Seob Shin
1,
Imene Arab
1,
Kyoungho Suk
2 and
Won-Ha Lee
1,*
1
School of Life Sciences, BK21 Plus KNU Creative BioResearch Group, Kyungpook National University, Daegu 41566, Republic of Korea
2
Department of Pharmacology, Brain Science and Engineering Institute, BK21 Plus KNU Biomedical Convergence Program, School of Medicine, Kyungpook National University, Daegu 41944, Republic of Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(5), 2670; https://doi.org/10.3390/ijms25052670
Submission received: 31 January 2024 / Revised: 20 February 2024 / Accepted: 22 February 2024 / Published: 25 February 2024

Abstract

:
Over the past century, molecular biology’s focus has transitioned from proteins to DNA, and now to RNA. Once considered merely a genetic information carrier, RNA is now recognized as both a vital element in early cellular life and a regulator in complex organisms. Long noncoding RNAs (lncRNAs), which are over 200 bases long but do not code for proteins, play roles in gene expression regulation and signal transduction by inducing epigenetic changes or interacting with various proteins and RNAs. These interactions exhibit a range of functions in various cell types, including macrophages. Notably, some macrophage lncRNAs influence the activation of NF-κB, a crucial transcription factor governing immune and inflammatory responses. Macrophage NF-κB is instrumental in the progression of various pathological conditions including sepsis, atherosclerosis, cancer, autoimmune disorders, and hypersensitivity. It orchestrates gene expression related to immune responses, inflammation, cell survival, and proliferation. Consequently, its malfunction is a key contributor to the onset and development of these diseases. This review aims to summarize the function of lncRNAs in regulating NF-κB activity in macrophage activation and inflammation, with a particular emphasis on their relevance to human diseases and their potential as therapeutic targets. The insights gained from studies on macrophage lncRNAs, as discussed in this review, could provide valuable knowledge for the development of treatments for various pathological conditions involving macrophages.

1. Introduction

Inflammation, a natural and complex biological response, is crucial in defending against infection [1,2]. Yet, chronic or uncontrolled inflammation can be harmful, contributing to the development of various diseases [3]. Diet, exercise, smoking, and stress, as lifestyle factors, can impact bodily inflammation levels, underscoring a healthy lifestyle’s importance in preventing and managing diseases [4]. In some instances, medical treatments target inflammation to ease symptoms and control disease progression [5,6].
Macrophages, innate immune cells with antigen-presenting capabilities, are central to inflammatory responses, significantly impacting acute and chronic inflammation [7]. Macrophages exhibit a dual role in inflammation; pro-inflammatory M1 macrophages drive inflammation, while anti-inflammatory M2 macrophages aid in tissue repair and inflammation resolution [8]. The balance between M1 and M2 macrophages is crucial, as their imbalance can result in immune-related diseases, including chronic inflammation, cancer, hypersensitivity, and autoimmune disorders [9]. This dichotomy is not strictly binary but represents a spectrum of activation states that macrophages can adopt in response to various signals within their microenvironment.
M1 macrophages, induced by IFN-γ alone or in combination with microbial products like lipopolysaccharide (LPS), are potent effectors of the pro-inflammatory response [10]. They produce high levels of pro-inflammatory cytokines (e.g., TNF-α, IL-1β, IL-6), promote Th1 responses, and are highly microbicidal. Their role is crucial in defending against intracellular bacteria and viruses, as well as in tumor suppression. However, the chronic activation of M1 macrophages is implicated in the pathogenesis of autoimmune and inflammatory diseases due to their tissue-damaging effects [11,12]. M2 macrophages are induced by IL-4 and IL-13 and are associated with tissue repair, resolution of inflammation, and immune regulation [5,10]. They produce anti-inflammatory cytokines (e.g., IL-10, TGF-β) and growth factors that promote tissue remodeling, wound healing, and angiogenesis. While essential for recovery from injury, excessive or inappropriate M2 activity can contribute to fibrosis, allergic responses, and the suppression of effective immune responses against tumors [11,12].
An imbalance between M1 and M2 macrophages constitutes a pivotal pathogenetic mechanism in numerous diseases. A predominance of M1 macrophages can lead to chronic inflammatory states, including rheumatoid arthritis, inflammatory bowel disease, and atherosclerosis. In these conditions, sustained inflammation results in tissue damage and drives disease progression [9,13]. Conversely, an excess of M2 macrophages can promote tumor growth and metastasis and contribute to fibrotic diseases and certain infections by failing to eliminate pathogens adequately [8,14].
Given their central role in both the promotion and resolution of inflammation, as well as in tissue repair, macrophages present a valuable target for therapeutic intervention. Strategies to modulate macrophage activity and polarization could offer new avenues for treating a wide range of diseases. For instance, therapies aimed at promoting M1 activity could enhance anti-tumor immunity or combat chronic infections, while strategies to boost M2 functions might aid in tissue repair and resolution of inflammation [8,15]. Rebalancing the M1/M2 ratio could provide therapeutic benefits in diseases characterized by an imbalance in macrophage polarization. Therefore, targeting macrophage activity represents a promising approach for the development of novel therapeutic strategies aimed at a variety of inflammatory and immune-related diseases.
The nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) is pivotal in regulating inflammation, particularly in macrophages. NF-κB, comprising p50, p52, p65 (RelA), RelB, and c-Rel subunits, activates genes related to immune response, inflammation, and cell survival by forming homodimers and heterodimers [16]. Activation of NF-κB involves the nuclear translocation of the p65/p50 heterodimer, whereas the p50 homodimer acts as an inhibitor [17]. NF-κB activation occurs via two primary pathways: the canonical (classical) and the non-canonical (alternative) [18]. Differing in activation mechanisms and stimuli, these pathways enable precise and dynamic NF-κB regulation. While the canonical pathway primarily mediates rapid and transient NF-κB activation in response to pro-inflammatory stimuli, the non-canonical pathway is involved in the regulation of adaptive immune responses, lymphoid organ development, and maintenance of immune homeostasis. Induced by cytokines like tumor necrosis factor (TNF)-α and interleukin (IL)-1, the canonical pathway involves TNF receptor-associated factor 6 (TRAF6) and the IKK complex (IKKα, IKKβ, IKKγ), leading to IκB phosphorylation. Subsequent degradation of IκB liberates p65/p50 heterodimer for nuclear translocation. Initiated by TNF receptor superfamily members like lymphotoxin β receptor (LTβR) or B cell-activating factor receptor (BAFF-R), the non-canonical pathway is then mediated by NF-κB-inducing kinase (NIK) and IKKα for phosphorylation and processing of p100 to generate p52. The resulting RelB/p52 heterodimers then translocate to the nucleus for transcriptional activation [18]. Dysregulation in either pathway can lead to numerous inflammatory and immune-related diseases.
Long noncoding RNAs (lncRNAs) constitute a heterogeneous group of RNA molecules exceeding 200 nucleotides in length that do not encode proteins, yet are pivotal in regulating a myriad of biological functions [19]. These molecules are characterized by their complex roles in chromatin organization, gene transcription, RNA splicing, and epigenetic modifications [20,21,22]. LncRNAs influence chromatin accessibility and gene transcriptional activity by directing chromatin-modifying enzymes to specific genomic regions (Figure 1). This action can either enhance or suppress gene expression, contingent upon the context and the nature of the modifications implemented [20,23]. The synthesis of lncRNAs mirrors that of mRNAs, encompassing transcription primarily by RNA polymerase II, capping, splicing, and polyadenylation [24]. Nonetheless, lncRNAs generally exhibit reduced splicing efficiency and a diminished rate of nuclear-to-cytoplasmic export relative to mRNAs. Their genomic distribution is varied, including intergenic regions, areas overlapping with, or antisense to, protein-coding genes [19]. This diversity implies a broad spectrum of mechanisms through which lncRNAs can modulate gene expression and cellular functionality.
In their mechanisms of action, lncRNAs engage in various cellular processes. They function as guides, recruiting chromatin-modifying enzymes to specific genomic locations, thereby reshaping the epigenetic landscape and directly influencing gene expression at the DNA level [23,25]. As decoys, lncRNAs can attract and sequester transcription factors or other proteins away from their DNA targets, indirectly impacting transcription [21,22,26]. Moreover, serving as scaffolds, lncRNAs facilitate the assembly of multiple proteins into ribonucleoprotein complexes, enabling the orchestration of intricate molecular operations. Additionally, lncRNAs partake in post-transcriptional regulation by interacting with signaling mediators and by absorbing microRNAs (miRNAs) [27]. Through miRNA sequestration, lncRNAs obstruct miRNA attachment to target mRNAs, thereby moderating post-transcriptional gene silencing and influencing mRNA stability and translation.
With the advancement of high-throughput sequencing and bioinformatics tools, a growing number of lncRNAs have been identified as key regulators in macrophage activity, particularly in the regulation of NF-κB signaling pathways. LncRNAs can modulate NF-κB activity through direct interaction with NF-κB itself or its signaling components, or by influencing the expression of genes that are part of the NF-κB signaling pathway. This regulation can have profound effects on the inflammatory response of macrophages, impacting the development and progression of a wide array of human diseases (Figure 2). This review aims to comprehensively explore the multifaceted roles of lncRNAs in modulating NF-κB activity within macrophages. By highlighting specific examples of lncRNAs that have been implicated in this process, it underscores their potential as biomarkers for disease diagnosis and prognosis, as well as targets for therapeutic intervention.

2. LncRNAs That Modulate Macrophage NF-κB Activity in Sepsis

Sepsis is a life-threatening condition marked by a dysregulated immune response to infection, leading to widespread inflammation, organ dysfunction, and potentially organ failure and death [28]. Often caused by bacteria, viruses, fungi, or other pathogens, a prime stimulant is LPS, an endotoxin found in the cell wall of Gram-negative bacteria [29]. The systemic inflammatory response, particularly post-bloodstream infection, triggers an excessive release of pro-inflammatory cytokines and chemokines. These mediators damage the endothelial cells lining the walls of blood vessels, resulting in increased permeability. This, in turn, leads to fluid leakage, tissue swelling, and subsequently edema [30]. Persistent inflammation can impact multiple organs and systems, with cytokines like TNF-α being key mediators of sepsis [31]. Commonly affected organs include the lungs, heart, kidneys, liver, and even the central nervous system.
NF-κB-mediated inflammation in macrophages, followed by M1 polarization, is pivotal in the pathogenesis of sepsis [32]. Pattern recognition receptors, including Toll-like receptors (TLRs), allow macrophages to identify pathogen-associated molecular patterns on invading microorganisms [29]. Activation of NF-κB, triggered by LPS interaction with TLR4 via the canonical pathway, upregulates numerous pro-inflammatory genes, such as cytokines, chemokines, adhesion molecules, and enzymes like inducible nitric oxide synthase [33]. NF-κB serves as a primary target for immunomodulatory therapies aimed at regulating inflammation and reestablishing immune equilibrium in many diseases, including sepsis. Such strategies might involve inhibiting NF-κB activation or employing targeted therapies to neutralize specific pro-inflammatory cytokines.
The lncRNA nuclear paraspeckle assembly transcript 1 (NEAT1), confined to the nucleus, is instrumental in forming paraspeckles, subnuclear structures involved in antiviral responses [34]. The levels of NEAT1 are increased by more than two-fold in the sera of sepsis patients [35,36,37,38]. Induced by LPS in the human monocytic leukemia cell line THP-1, NEAT1 enhances inflammatory responses by sponging miR-17-5p, thereby stabilizing TLR4 mRNA (the miR-17-5p/TLR4 axis) (Figure 3) [38]. In Kupffer cells and the murine macrophages, LPS-induced NEAT1 promotes inflammatory activities via the Let-7q/TLR4 axis [35,36,37,39]. Additionally, NEAT1 also facilitates M1 polarization in macrophages through the miR-125a-5p/TRAF6/TGF-β-activated kinase 1 (TAK1) axis [40], underscoring its potential as a therapeutic target for sepsis.
Metastasis-associated lung adenocarcinoma transcript 1 (MALAT1), a multifunctional lncRNAs in macrophages, is extensively studied. In late-onset sepsis patients, increased blood MALAT1 levels correlate with disease severity [41]. The expression of MALAT1 increases in activated macrophages, exhibiting an elevation greater than two-fold in mouse peripheral blood mononuclear cells (PBMCs) and more than six-fold in THP-1-derived macrophages, especially following LPS treatment [41,42]. Animal studies show that MALAT1 expression surges with sepsis induction; reducing MALAT1 lessens inflammation and mortality [41,43,44,45], possibly by suppressing M1 and enhancing M2 macrophage polarization [45]. In an LPS-induced septic lung injury model, intravenous MALAT1-specific small interfering RNA (siRNA) decreases inflammatory cytokines and immune cells in bronchoalveolar lavage fluid by inhibiting the p38 mitogen-activated protein kinase (MAPK)/p65 pathway [44].
However, several reports have shown contradictory evidence, indicating a significant decrease in MALAT1 expression accompanied by an increase in hsa-miR-346 levels in patients with sepsis. Activated human and mouse macrophages downregulate MALAT1 expression in an NF-κB-dependent manner [42,46]. MALAT1 interacts with NF-κB, inhibiting its DNA-binding activity and, consequently, the expression of pro-inflammatory cytokines [42]. Furthermore, MALAT1 modulates macrophage proliferation by sequestering hsa-miR-346, thereby stabilizing the mRNA of small mothers against decapentaplegic homolog 3 (SMAD3). SMAD3 is a receptor-regulated signaling adaptor that is activated by serine kinases. These conflicting findings underscore the need for future research on MALAT1’s role in sepsis.
In sepsis patients, the levels of long noncoding RNA plasmacytoma variant translocation 1 (PVT1) are elevated, showing an increase of more than two-fold compared to the healthy control group. This elevation correlates with increased pro-inflammatory mediators and survival rates. [47,48]. LPS induces PVT1 expression in THP-1 cells, which in turn amplifies NF-κB activity via p38 stimulation [49]. Elevated PVT1 expression, promoting M1 polarization through the miR-29a/high-mobility group box 1 (HMGB1) axis, is observed in heart-infiltrating macrophages of septic mice [50]. HMGB1, released from the cells, can activate TLR4 in both autocrine and paracrine manners [51]. PVT1 is also highly expressed in osteoarthritis patients’ serum and in the LPS-stimulated C28/12 chondrocyte cell line, activating the TLR4/NF-κB pathway via the miR-93-5p/HMGB1 axis [52]. Additionally, PVT1 levels rise in myocardial tissues and heart-infiltrating macrophages during sepsis-induced myocardial injury [50].
The expression level of lncRNA MEG3 is significantly reduced in patients with sepsis, and this reduction has prognostic significance [53]. In macrophages, MEG3 overexpression inhibits LPS-induced apoptosis by downregulating BAX and upregulating Bcl-2. It also suppresses inflammatory factor expression by inhibiting NF-κB signaling [53]. This suggests that the reduced MEG3 expression may exacerbate sepsis by increasing inflammation and inhibiting apoptosis in macrophages. Further research is needed to elucidate MEG3’s role in sepsis.
In sepsis patients, the lncRNA colorectal neoplasia differentially expressed (CRNDE) exhibits elevated expression in peripheral blood, with higher levels correlating to improved survival rates [54]. CRNDE intensifies LPS-induced NF-κB activation and subsequent pro-inflammatory cytokine release in THP-1 cells via the miR-181-5p/TLR4 axis [54].
These reports highlight the intricate relationship between lncRNAs and NF-κB in the context of sepsis, impacting inflammatory activation and macrophage polarization. The influence of these lncRNAs on cytokine release, cell polarization, and apoptosis is notable, and their varied expression in sepsis patients suggests potential as biomarkers. Targeting these lncRNAs to regulate NF-κB activation offers promising avenues for immunomodulatory therapies to manage inflammation and restore immune balance in sepsis. However, the contrasting roles of specific lncRNAs, like MALAT1, necessitate further research. A deeper understanding of these lncRNAs’ roles could lead to innovative diagnostic and therapeutic strategies, improving management and outcomes in sepsis.

3. LncRNAs That Modulate Macrophage NF-κB Activity in Atherosclerosis

Macrophages are central in atherosclerosis development, marked by arterial plaque build-up. The process initiates with low-density lipoprotein (LDL) cholesterol accumulation in arterial walls, undergoing oxidation and eliciting inflammation [55,56]. Modified LDL attracts monocytes from blood, transforming into macrophages in the arterial wall. These macrophages consume oxidized LDL (oxLDL), forming lipid-laden foam cells and creating fatty streaks, early atherosclerosis signs [57,58]. M1 macrophages exacerbate inflammation by releasing cytokines, attracting more immune cells [59,60]. Chronic inflammation leads to fibrous cap formation over plaques and extracellular matrix accumulation. Macrophages also degrade this matrix, heightening plaque instability and increasing heart attack and stroke risks [57,61]. Macrophages can also contribute to the resolution of inflammation and healing processes [62]. In atherosclerosis, inflammation resolution is overshadowed by ongoing inflammation and plaque growth.
NF-κB, activated by stimuli such as oxidative stress, cytokines, and oxLDL, exacerbate atherosclerosis by promoting lipoprotein uptake, foam cell formation, and attracting more immune cells [57,63]. This activation also destabilizes plaques by encouraging matrix metalloproteinase (MMP) secretion, increasing plaque rupture risks [64]. Chronic NF-κB activation sustains the inflammation characteristics of advanced atherosclerosis in conditions like coronary artery diseases (CADs) and myocardial infarction (MI) [65]. Given its crucial role in macrophage activation atherosclerosis progression, NF-κB presents a potential target for therapies aimed at reducing inflammation and slowing atherosclerosis progression [66].
Increased NEAT1 expression levels in the PBMCs and sera of atherosclerosis patients have been noted [67,68]. The expression level of NEAT1 was found to be increased by more than two-fold in PBMCs of CAD patients [67]. NEAT1, induced by oxLDL in THP-1 cells, contributes to pro-inflammatory responses by enhancing p65 phosphorylation, followed by paraspeckle formation [69,70]. It is also induced in bone marrow-derived macrophages (BMDMs) treated with titanium particles and promotes NF-κB activation, NLRP3 inflammasome formation, and M1 polarization via the miR-188-5p/Bruton’s tyrosine kinase (BTK) axis [71]. NEAT1 also stimulates pro-inflammatory cytokine and reactive oxygen species (ROS) production and subsequent foam cell formation by sponging miR-342-3p in THP-1 cells [70] or miR-128 in the murine macrophage-like cell line RAW264.7 [72]. These reports agree with NEAT1 being expressed in activated macrophages and enhancing pro-inflammatory changes. One contradicting study, however, reported decreased NEAT1 in the PBMCs of post-MI patients and enhanced macrophage inflammation in NEAT1-knockout mice (Table 1) [73].
Elevated lncRNA PVT1 levels have been detected in the serum of atherosclerosis patients [74]. Inhibiting PVT1 in animal models reduces atherosclerotic plaques by increasing HDL levels and suppressing the MAPK/NF-κB pathway and pro-atherogenic factors [74]. In serum samples of atherosclerosis patients and during oxLDL-induced THP-1 cell foam cell differentiation, there is a notable increase in lncRNA small nucleolar RNA host gene (SNHG)16 and a decrease in miR-17-5p [75]. SNHG16 amplifies macrophage proliferation and pro-inflammatory responses in atherosclerosis through the miR-17-5p/NF-κB axis [75].
LncRNA X-inactive specific transcript (XIST), known for its role in X-chromosome inactivation, has been found to be elevated more than two-fold in the serum of atherosclerosis patients, oxLDL-treated vascular smooth muscle cells, and the U937 human monocytic leukemia cell line [76]. XIST influences atherosclerosis by promoting proliferation and inhibiting apoptotic cell death through the miR-599/TLR4 axis [76]. This finding aligns with other studies that show that apoptosis inhibition aggravates atherogenesis by increasing macrophage proliferation and plaque formation [77,78].
LncRNA H19 is found at elevated levels in the serum of atherosclerosis patients [79,80,81,82]. OxLDL stimulates H19 expression in macrophages [83], aorta vascular smooth muscle cells [79,80], and human umbilical vein endothelial cells (HUVECs) [84]. In macrophages, H19 augments oxLDL-induced lipid accumulation, ROS generation, and NF-κB activation [83,85]. Similarly, in HUVECs, H19 heightens NF-κB activation by increasing p38 and p65 activity [86]. These findings suggest that H19 could be a promising therapeutic target for atherosclerosis treatment.
In atherosclerosis and CAD patients, MALAT1 levels rise more than two-fold and subsequently fall after treatment [87,88,89]. MALAT1 impacts various macrophage processes like foam cell formation, autophagy, and pyroptosis [90,91,92]. OxLDL prompts NF-κB-dependent MALAT1 expression in THP-1 cells. MALAT1 then enhances lipid uptake and foam cell formation by promoting scavenger receptor CD36 expression [88,90,91]. MALAT1 also enhances NF-κB activation and subsequent inflammation by sponging miR-330-5p [91]. Further, oxLDL-induced autophagy in macrophage is mediated by MALAT1, which activates the MAPK/NF-κB pathway and inhibits sirtuin 1 (SIRT1), a key transcription factor deacetylase [92,93]. NLRP3 inflammasome-mediated pyroptosis, a programmed cell death as a defense mechanism against intracellular pathogens, is also influenced by MALAT1 [94]. In diabetic atherosclerosis models, a cinnamic acid derivative reduces inflammasome activation and pyroptosis by suppressing MALAT1 [95]. Extracellular vesicles (EVs) such as exosomes are crucial for cell-to-cell communication, transferring proteins and lncRNAs [96]. M1 macrophages have been found to release MALAT1-containing EVs, which regulate myocyte proliferation and angiogenesis in MI models [97]. These findings underscore MALAT1’s role in atherosclerosis: it is upregulated in activated macrophages and influences various processes including lipid uptake, foam cell formation, and cell death.
However, contrary reports exist regarding the role of MALAT1 in atherosclerosis. It was observed that in atherosclerosis patients and oxLDL-treated THP-1 cells, MALAT1 levels decrease [88]. Reduced MALAT1 leads to increased lipid and total cholesterol accumulation in THP-1 cells via the miR-17-5p/ATP-binding cassette subfamily A member 1 (ABCA1) axis [88]. ABCA1 is known to facilitate cholesterol efflux, thereby reducing foam cell formation [98]. Additionally, MALAT1 deficiency in certain mouse models has been linked to accelerated macrophage inflammation and atherosclerosis [99]. Exosomal MALAT1 from oxLDL-treated HUVECs promotes a transition from M1 to M2 macrophages [100]. These findings suggest potential anti-atherogenic properties of MALAT1, highlighting the need for further research to clarify its role in atherosclerosis.
Notably, the expression levels of lncRNA HOX transcript antisense intergenic RNA (HOTAIR) are decreased in the peripheral blood lymphocytes of atherosclerosis patients and oxLDL-treated RAW264.7 cells [101]. HOTAIR overexpression reduces pro-inflammatory cytokine expression while boosting anti-inflammatory cytokines, achieved by inhibiting NF-κB activity. This suppression occurs through HOTAIR’s enhancement of fragile X-related protein 1 (FXR1) levels, a protein moving between the nucleus and cytoplasm and associating with polyribosomes [101,102].
These reports underscore the complex relationship between various lncRNAs and macrophage NF-κB in atherosclerosis. These lncRNAs impact crucial aspects such as lipid uptake, foam cell formation, inflammation, and cell death in macrophages. Given their link to NF-κB activation, targeting these lncRNAs for NF-κB modulation presents a promising approach to managing atherosclerosis by restoring immune equilibrium and curbing inflammatory activation. It is intriguing that certain lncRNAs, such as MALAT1 and HOTAIR, have been identified to play conflicting roles in atherosclerosis. Variations in the stages of atherosclerosis or CAD examined, the measurement techniques utilized, environmental factors, or the experimental model systems employed could account for these discrepancies. Alternatively, the overall impact of these lncRNAs may differ based on the dominant signaling pathways activated in particular contexts or disease states. Despite the conflict, their significant influence on macrophage function and disease progression is evident. Further research is essential to unravel the full potential of these lncRNAs in atherosclerosis treatment.
Table 1. LncRNAs regulating macrophage NF-κB activity in diseases. LncRNAs that enhance NF-κB activity are indicated in red, while those that suppress NF-κB activity are shown in blue.
Table 1. LncRNAs regulating macrophage NF-κB activity in diseases. LncRNAs that enhance NF-κB activity are indicated in red, while those that suppress NF-κB activity are shown in blue.
LncRNADiseaseTargetFunctionRefs.
Group 1. LncRNAs directly affecting NF-κB activity
CARLRCeliac diseasep65LPS-induced, regulates p65 translocation, promotes pro-inflammatory cytokine production[103]
COX2Neuroinflammationp65LPS-induced, promotes p65 translocation, enhances inflammasome formation, suppresses autophagy[104]
COX2SLESWI/SNF/NF-κBLPS-induced, enhances the expression of NF-κB-induced late inflammatory genes[105]
MALAT1Sepsisp65/p50LPS-induced, retrains NF-κB from promoter, regulates pro-inflammatory cytokine expression[42]
PACERCancerp50LPS-induced, sequesters p50 and promotes p300-mediated histone acetylation, enhances COX2 expression[106]
PINTCancerp65 and EZH2LPS-induced, bridges p65 and EZH2, activates TNFα transcription[107]
Group 2. LncRNAs affecting pathways that regulate NF-κB activity
CHRFCancermiR-489/MyD88Silica-induced, promotes inflammatory responses and fibrosis[108]
CRNDECancermiR-181a-5p/TLR4LPS-induced, promotes inflammatory responses. Also increased in AML and IgA nephropathy[54,109]
MALAT1AtherosclerosisSIRT1/MAPKox-LDL-induced, promotes autophagy, reduces apoptosis[93]
MIR222HGAllergic rhinitismiR-146a-5p/TRAF6Deceased in patients, causing the dominance of type 2 response. Promotes M1 and suppresses M2 polarization[110]
NAILUlcerative colitisWip1LPS-induced, promotes p65 phosphorylation by blocking Wip1 action, increases inflammatory response[111]
NEAT1OsteolysismiR-188-5p/BTK
KLF/BTK
Induced by titanium particles, activates inflammasome and NF-κB, enhances M1 polarization[71]
NEAT1SepsismiR-17-5p/TLR4LPS-induced, promotes inflammatory response by stabilizing TLR4 mRNA[38]
NEAT1Sepsislet-7a/TLR4LPS-induced, promotes inflammatory response by stabilizing TLR4 mRNA[35]
NKILAAsthmaIκBAnti-inflammation by inhibiting IκB phosphorylation, induces M2 polarization[112,113]
PVT1SepsismiR-29a/HMGB1/TLR4LPS-induced, promotes inflammation and M1 polarization. Also increased in osteoarthritis patients[50,52]
SNHG1Cancer/SepsisHMGB1/TLR4Enhances TLR4 signaling by interaction with HMGB1, promotes M1 polarization[114]
XISTAtherosclerosismiR-599/TLR4OxLDL-induced in macrophages/vascular smooth muscle cells, promotes proliferation and suppresses apoptosis[76]
Group 3. LncRNAs affecting NF-κB activity with unknown mechanism of action
DCST1-AS1Cancerp65Activates NF-κB signaling pathway in both cancer cells and macrophages. Promotes M2 polarization[115]
FTXCirrhosisNF-κBAnti-inflammatory function. Decreased in patients, causing enhancement in inflammation[116]
HOTAIRCancerIκBαLPS-induced, activates NF-κB by degrading IκB, regulates metabolic reprogramming by inducing GLUT1[117]
HOTAIRAtherosclerosisFXR1Anti-inflammation by inactivating NF-κB via FXR1. Decreased in patients and oxLDL-treated macrophages.[101,102]
NEAT1Atherosclerosisp65, ERKox-LDL-induced, regulates p65 and ERK phosphorylation, promotes TNF-α secretion.[34,69]
MEG3Sepsisp65Decreased in sepsis patients. LPS-induced, inhibits p65 phosphorylation. Downregulates inflammation and apoptosis.[53]
SNHG16AtherosclerosismiR-17-5pox-LDL-induced, promotes proliferation and inflammatory response. Also increased in cancer and diabetes[75,118,119]
Group 4. LncRNAs transferred to macrophages via exosomes or EVs
FGD5-AS1CancerP300/STAT3/NF-κBContained in exosomes from pancreatic cancer cells, activates STAT3/NF-κB, promotes M2 polarization[120]
AP000439.2CancerSTAT3, P65Contained in exosomes from cancer cells, promotes macrophage M2 polarization[121]
GAS5Allergic rhinitismTORC1/ULK1/ATG13Activates NF-κB and promotes M1 polarization by suppressing autophagy-dependent degradation of IKKa/b[122]
HOTTIPCancer-Contained in M1-derived exosomes. Suppresses cancer growth via TLR5 activation. Promotes M1 polarization[123]
MALAT1Acute pancreatitismiR-181a-5p
/HNGB1/TLR4
Carried by EVs originating from pancreatic cancer cells, promotes M1 polarization[124]

4. LncRNAs That Modulate Macrophage NF-κB Activity in Cancer

The role of macrophage inflammation in cancer is multifaceted and contradictory. M1 macrophages, typically anti-tumorigenic, can attack tumor cells and stimulate immune responses. Conversely, M2 macrophages often aid tumor growth by supporting angiogenesis, suppressing immune responses, and facilitating tissue remodeling [125]. Generally, tumor-associated macrophages (TAMs) exhibit an M2 phenotype, supporting tumor growth and metastasis and contributing to an immunosuppressive tumor environment [126,127]. Given their significant impact on cancer progression, TAMs are being investigated as therapeutic targets, with strategies focusing on inhibiting their tumor-promoting functions or reprogramming them to combat tumors.
The activation of NF-κB in macrophages plays a crucial role in cancer development and progression. In TAMs, NF-κB activation leads to the production of cytokines, growth factors, and enzymes that promote tumor growth and suppress anti-tumor immune responses [128,129]. NF-κB can also alter the immune microenvironment, potentially inducing immune checkpoint molecules that weaken the immune response against tumors [130]. Additionally, NF-κB-activated macrophages can produce angiogenic factors, aiding tumor vascularization [131,132,133]. They can also stimulate matrix metalloproteinases (MMPs), breaking down extracellular matrix barriers and facilitating cancer cell spread [134,135]. Thus, macrophage NF-κB is implicated in various aspects of cancer progression and targeting macrophage NF-κB has emerged as a prominent focus in cancer treatment strategies [18].
LncRNA DC-STAMP domain containing 1-antisense 1 (DCST1-AS1) has been investigated in various cancers, including gastric, colorectal, cervical, breast, glioblastoma, endometrial, and HCC [136,137,138,139,140,141,142]. In these cancers, increased DCST1-AS1 expression correlates with larger tumors and shorter survival and DCST1-AS1 promotes cancer cell proliferation and metastasis, and inhibits apoptosis, by sponging miRNAs [136,137,138,139,140,141,142]. Notably, in oral squamous cell carcinoma, DCST1-AS1 advances tumor progression by enhancing NF-κB activity in cancer cells and macrophages [115]. The expression of DCST1-AS1 showed a more than three-fold increase in oral squamous cell carcinoma cells compared to normal cells [115]. Elevated DCST1-AS1 in cancer cells and M2 macrophages is linked to tumor growth and cancer cell proliferation. NF-κB antagonists revealed that DCST1-AS1 enhances cancer progression and M2 macrophage polarization through NF-κB-mediated mechanisms [115].
The lncRNA FGD5 antisense RNA 1 (FGD5-AS1) shows elevated levels in non-small-cell lung cancer and pancreatic cancer, correlating with metastasis and poor prognosis [120,143]. FGD5-AS1-containing exosomes from these cancers induce M2 macrophage polarization [120]. FGD5-AS1 links acetyltransferase p300, STAT3, and NF-κB, leading to acetylated STAT3/p65 complex and transcriptional activation [120,144]. STATs are crucial transcription factors in macrophage polarization, with STAT1 being integral to M1, and STAT3/6 to M2 polarization. [145]. In cervical cancer, FGD5-AS1, via the miR-129-5p/bone marrow stromal cell antigen 2 (BST2) axis, promotes tumor growth and M2 polarization [146]. BST2, a lipid raft-associated protein, is implicated in cell proliferation and immune response [147,148]. Collectively, FGD5-AS1 augments tumor growth by enhancing cancer progression and M2 macrophage polarization.
The lncRNA AP000439.2 has recently been identified as a prognostic marker for renal cell carcinoma (RCC) patient survival [149,150]. Exosomes from human RCC cell lines have been shown to induce M2 polarization in co-cultured THP-1 cells [121]. AP000439.2 promotes M2 polarization through the phosphorylation of STAT3 and the NF-κB p65 subunit, which, in turn, enhances the migration potential of cocultured cancer cell lines. The impact of exosomal AP000439.2 on macrophage M2 polarization and RCC growth has been confirmed in a xenograft tumor mouse model [121].
LncRNA Five Prime to Xist (FTX), an evolutionarily conserved regulator of XIST expression, is associated with various conditions including malignancies, endometriosis, and stroke, functioning through miRNA sponging [151,152]. Liu et al. observed decreased FTX levels in cirrhosis patients, linking them to abnormal activation of CD14+ CD16+ monocytes via the miR-545/T cell immunoglobulin and mucin domain 3 (Tim-3) axis [116,153]. Moreover, FTX suppression in THP-1 cells increases NF-κB activity and pro-inflammatory cytokine expression, suggesting that a reduction in FTX might accelerate tumor progression by enhancing inflammation in the tumor microenvironment (TME) [116].
LncRNA HOTAIR, known for its role in gene regulation and epigenetic modifications, is implicated in various human diseases [154]. It is often overexpressed in cancer, contributing to tumor progression, metastasis, and poor prognosis by altering gene expression related to the cell cycle, apoptosis, and metastasis [154,155]. HOTAIR is also associated with central nervous system disorders, fibrosis, and inflammatory conditions, impacting cellular processes and immune responses [156,157,158,159]. It regulates glucose transporter isoform 1 (GLUT1) expression in human neuroblastoma cells and macrophages by stimulating NF-κB activity, suggesting a role in metabolic reprogramming in cancer [117,160]. In addition, inflammatory activation of macrophages triggers HOTAIR expression, which then promotes NF-κB activation and cytokine gene expression by aiding in the degradation of IκBα [117]. HOTAIR’s expression pattern in cancer tissue macrophages remains unexplored and warrants future investigation.
Elevated levels of lncRNA HOXA transcript at the distal tip (HOTTIP) have been observed in AML patients and cell lines, such as U937 and THP-1 [161]. HOTTIP facilitates cell proliferation via the miR-608/DET1- and DDB1-associated 1 (DDA1) axis, with DDA1 being a gene known for its oncogenic properties [161,162]. In squamous cell carcinoma, M1-derived exosomes containing HOTTIP inhibit cancer cell proliferation and induce apoptosis by activating the TLR5/NF-κB pathway [123]. Additionally, exosomal HOTTIP influences the M1 polarization of circulating monocytes [123]. The comprehensive role of HOTTIP in cancer progression remains an area for future exploration.
Cyclooxygenase (COX)2, linked with inflammation in immune cells, is implicated in several cancers [163]. LncRNA p50-associated COX2 extragenic RNA (PACER), located upstream of the COX2 promoter, regulates COX2 expression [106]. PACER, through its association with p50, facilitates p65/p50 heterodimer binding to the COX2 promoter, recruiting p300 histone acetyltransferase [106]. Its expression is upregulated in various cancer tissues, influencing COX2 and PGE2 synthesis and cancer cell proliferation, migration, and invasion [164,165,166].
LncRNA cardiac hypertrophy-related factor (CHRF) functions as an oncogene, promoting migration and invasion in various tumor types [167]. In a silica-induced pulmonary fibrosis mouse model, CHRF activates inflammatory and fibrotic pathways via the miR-489/MyD88 and miR-489/SMAD3 axes, with SMAD3 being an adaptor in receptor-regulated signaling [108,168]. CHRF’s pro-inflammatory effects are also observed in LPS-induced acute lung injury [169]. However, its specific role in macrophage inflammation within the TME remains unclear, necessitating further research.
LncRNA SNHG1, commonly overexpressed in various cancers as an oncogene, affects cellular signaling via interactions with miRNAs and signaling regulators [170]. In cholangiocarcinoma cell lines, SNHG1 is associated with increased proliferation and invasion, mediated by NF-κB activation through the miR-140/TLR4 axis, contributing to an inflammatory TME [171]. In an LPS-induced acute lung injury model, SNHG1 is upregulated in M1 polarized THP-1 cells, enhancing NF-κB activation and inflammation through interaction with HMGB1 [114]. However, SNHG1’s specific role in macrophage-related TME remains unexplored.
These findings highlight the intricate relationship between macrophages, lncRNAs, and NF-κB in cancer, affecting cell proliferation, invasion, inflammation, and macrophage polarization. The dichotomy of macrophages, especially TAMs, underscores their potential as therapeutic targets. Their influence extends to inflammation, TME modulation, angiogenesis, and immunosuppression, making them key in the interplay between cancer cells and the immune system. Understanding lncRNA-driven macrophage NF-κB regulation is essential for developing targeted cancer therapies. Despite advances, many aspects of lncRNA functions in cancer and inflammation require further exploration, presenting exciting opportunities for future research and potential therapeutic interventions.

5. LncRNAs That Modulate Macrophage NF-κB Activity in Autoimmunity and Hypersensitivity

NF-κB activation in macrophages plays a pivotal role in the pathogenesis of autoimmunity and hypersensitivity by orchestrating a series of inflammatory responses that can lead to tissue damage and disease progression [172]. In autoimmunity, aberrant NF-κB activation in macrophages contributes to the production of pro-inflammatory cytokines and chemokines, perpetuating a state of chronic inflammation and autoantibody production [173]. This not only disrupts immune tolerance but also promotes the survival and proliferation of autoreactive immune cells. In hypersensitivity reactions, NF-κB activation is crucial for the amplification of immune responses to allergens, leading to exaggerated inflammatory processes [174]. Such activation enhances the expression of molecules involved in antigen presentation and co-stimulation, facilitating the development of Th2 cell responses and the subsequent production of IgE [175]. By driving the inflammatory milieu in both autoimmunity and hypersensitivity, NF-κB activation in macrophages is central to the dysregulation of immune homeostasis and the pathophysiology of these conditions, highlighting its potential as a therapeutic target to modulate immune responses.
A more than two-fold increase in GAS5 levels has been found in exosomes from the nasal mucus of allergic rhinitis patients and a mouse model, promoting macrophage NF-κB activation and M1 polarization [122,176]. GAS5 influences this by inhibiting mTORC1/ULK1/ATG13-mediated autophagy, leading to NF-κB activation [122]. Specifically, GAS5’s activation of NF-κB signaling occurs through the suppression of autophagy-dependent degradation of IKKα/β in macrophages [122]. GAS5 also enhances M1 macrophage polarization, observed in hyperglycemia-induced differentiation in diabetics and in pneumonia-affected children’s macrophages [177,178]. It suppresses the Janus kinase 2/STAT3 pathway, promoting M1 while inhibiting M2 polarization via the miR-455-5p/SOCS3 axis [178]. In diabetes, hyperglycemia-induced GAS5 expression shifts polarization from M2 to M1 by upregulating STAT1 [177]. This shift is significant since M1 macrophage-driven inflammation worsens diabetic wounds, suggesting that reducing GAS5 could facilitate wound healing by promoting an M1-to-M2 transition. On the other hand, microglia exhibit increased GAS5 expression in multiple sclerosis and experimental autoimmune encephalomyelitis (EAE) mouse models [179]. GAS5 hinders M2 polarization by recruiting the polycomb repressive complex 2 (PRC2), suppressing topoisomerase-related function 4 transcription, crucial for M2 polarization [179]. Ito et al. found that reducing GAS5 expression leads to M2b polarization, while increasing it has the opposite effect [180]. GAS5 also plays a role in regulating NLRP3 inflammasome formation in cardiomyocytes [181]. In contrast, GAS5 expression levels are decreased in various cancer tissues [182,183]. This decrease is attributed to GAS5’s ability to induce apoptosis and inhibit tumor proliferation and metastasis through interactions with miRNAs and proteins. The M1-enhancing property of GAS5 might contribute to its reduced expression in cancer tissues, necessitating further investigation. Overall, research indicates that GAS5 boosts NF-κB-mediated inflammation and shifts macrophage polarization towards M1, while inhibiting GAS5 has the opposite effect in various diseases.
LncRNA-Cox2, located near the COX2 gene [184], is linked with chronic inflammatory diseases and innate immune activation [185,186]. In systemic lupus erythematosus (SLE) patients, serum lncRNA-Cox2 level was elevated more than three-fold and correlated with neurological symptoms [187]. It is highly induced in an NF-κB-dependent manner in activated macrophages and interacts with the switch/sucrose nonfermentable (SWI/SNF) chromatin-remodeling complex [184,188]. This interaction enhances the expression of NF-κB-induced late inflammatory genes [105]. In neuroinflammation, lncRNA-Cox2 promotes NLRP3-inflammasome formation and hinders autophagy by binding and facilitating the p65 subunit’s nuclear transport [104]. However, its role in inflammation is subject to debate, as it appears to suppress inflammation in M1 macrophages and BMDMs in septic mouse models [188,189], indicating a complex influence in inflammatory contexts.
During allergic disease development, a type 2 immune response led by Th2 cells, eosinophils, basophils, and M2 macrophages is predominant [190]. In allergic rhinitis clinical samples and animal models, a downregulation of the lncRNA miR222 host gene (MIR222HG) was observed, linking it to the type 2 response [110]. MIR222HG is upregulated in M1 macrophages but downregulated in M2 macrophages, with its overexpression reducing M2 polarization and allergic inflammation. MIR222HG activates the NF-κB pathway via the miR-146a-5p/TRAF6 axis [110]. This suggests that MIR222HG downregulation in allergic rhinitis contributes to pathogenesis by favoring M2 macrophage polarization, and its overexpression might alleviate the condition.
In contrast to MIR222HG, NF-κB interacting lncRNA (NKILA) directly promotes M2 macrophage polarization by suppressing NF-κB activation in asthma [112]. NKILA has been found to have anti-inflammatory effects and dysregulation of its expression has been observed in cancers as well as immune-related disorders such as epilepsy, osteoarthritis, periodontitis, coronary artery disease, and asthma [112,191]. Overexpression of NKILA alleviated airway hyperresponsiveness and asthmatic mice and reduced macrophage abundance through inhibition of the NF-κB pathway [112]. The inhibitory effect of NKILA on the NF-κB pathway was reported to be through the inhibition of IκB phosphorylation in laryngeal cancer cells.
Celiac disease, characterized by constant NF-κB activation, shows increased cardiac and apoptosis-related lncRNA (CARLR) expression in patient samples [103,192]. CARLR expression rises in THP-1 cells post-LPS stimulation. CARLR interacts with activated NF-κB following IκB dissociation, boosting the expression of cytokines such as IL-1β and COX2, also elevated in celiac disease samples [103]. This suggests a crucial role of CARLR in the disease’s inflammatory response.
NAIL, an evolutionarily conserved lncRNA, exhibits approximately a three-fold increase in expression in samples from patients with inflamed ulcerative colitis compared to those from non-inflamed cases [111]. It activates NF-κB by binding to Wip1 phosphatase, thereby removing Wip1 from its substrates, including p38 MAPK and p65, which play a key role in the inflammatory process of ulcerative colitis.
LncRNA P53-induced transcript (PINT) is upregulated in the intestinal mucosa of ulcerative colitis patients and in inflammatory bowel disease mouse models [107]. In macrophages, LPS induces PINT expression in an NF-κB-dependent manner. Acting as a scaffold, PINT enables the binding of p65 and EZH2, a histone methyltransferase and epigenetic regulator, to NF-κB sites on target promoters [193]. This complex stimulates TNF-α expression while inhibiting other pro-inflammatory mediators like CCL2/7 and intercellular adhesion molecule-1.
MALAT1, with high expression in severe acute pancreatitis patients’ plasma and corresponding mouse models, plays a critical role in this condition [124,194]. Its suppression reduces tissue injury and inflammation. Additionally, MALAT1 is found in serum EVs of acute pancreatitis patients, stimulating macrophages to activate TLR4/NF-κB signaling and M1 polarization via the miR-181a-5p/HMGB1 axis [124,194]. In endothelial progenitor cell-derived exosomes, MALAT1 triggers BMDM differentiation into osteoclasts, binding to miR-124 [195]. High glucose levels in macrophages also prompt MALAT1-containing exosome secretion [196], highlighting its role in cellular communication through EVs.
These studies underscore the pivotal roles of various lncRNAs in autoimmune, hypersensitivity, and inflammatory conditions. These lncRNAs modulate NF-κB activity, either directly or indirectly, impacting macrophage polarization and immune responses. The intricate relationship between these lncRNAs and macrophage NF-κB dynamics opens avenues for targeted therapies in such diseases. Given NF-κB’s association with inflammation, macrophage polarization, apoptosis, and pyroptosis, modulating lncRNA functions could favorably alter disease progression by influencing NF-κB activity.

6. Discussion

Research targeting lncRNAs in disease treatment shows promise, particularly through suppression or overexpression in animal models. For instance, suppressing MALAT1 with siRNA mitigates sepsis-related inflammation [43,44,45], while its overexpression exacerbates atherosclerosis severity [91]. Similarly, altering the M1/M2 macrophage polarization balance can influence disease progression. Knockdown of GAS5 in diabetic wound healing models encourages M1-to-M2 transition, enhancing wound healing [177], and its suppression reduces EAE progression by inhibiting M1 polarization [179]. Reducing lncRNA-Cox2 in HCC models strengthens M2 macrophage polarization, promoting tumor growth [197].
LncRNA-targeting therapies primarily use nucleic acid-based methods like antisense oligonucleotides (ASOs), RNA interference with siRNA or shRNA, and innovative approaches like CRISPR/Cas and exosome-mediated transfer [198,199]. While clinical trials have mainly focused on miRNAs, the exploration of lncRNAs as diagnostic markers and therapeutic targets is growing. For example, MALAT1 and lncRNA prostate cancer antigen 3 are being studied as diagnostic markers for prostate cancer. Trials involving ASncmtRNA-targeting ASOs, such as Andes-1537 [200], are assessing safety and efficacy in various cancers, indicating a progressing field in lncRNA-based treatments.
Contrary to its well-documented role in promoting inflammation, emerging evidence suggests that NF-κB may also be involved in anti-inflammatory responses within macrophages, highlighting the complex role of NF-κB in immune regulation [201]. This dual functionality depends on the context, including the specific cell type, the nature of the stimuli, and the timing of NF-κB activation. For example, many negative regulators of TLR signaling such as A20, IκB, and protein–tyrosine phosphatases are direct targets of NF-κB signaling [201]. SREBP1, a transcription factor influenced by NF-κB, contributes to the resolution of pro-inflammatory TLR4 signaling by reprogramming fatty acid metabolism in macrophages. This reprogramming leads to the production of anti-inflammatory fatty acids, further indicating a mechanism through which NF-κB may support anti-inflammatory responses [202]. These anti-inflammatory actions of macrophage NF-κB may complicate future efforts to manipulate its activity for disease treatment, necessitating careful design in therapeutic approaches.
The expression of lncRNAs that regulate macrophage NF-κB activity is implicated in various human diseases. Animal studies targeting these lncRNAs have yielded promising results, positioning them as potential future therapeutic targets. The primary focus will be on reducing macrophage inflammation in chronic inflammatory diseases, hypersensitivity, autoimmune disorders, and cancer. A secondary objective will involve balancing M1 and M2 macrophage polarization in these conditions. While the use of NF-κB-targeting lncRNAs as therapeutic agents is still in its early stages, the critical role of macrophage NF-κB in disease pathogenesis makes these targets particularly promising.

7. Conclusions and Future Directions

This review has comprehensively explored the dynamic interplay among lncRNAs, NF-κB activation, and macrophage function across various diseases. It elucidates how lncRNAs and NF-κB are mutually regulatory within a complex network that is pivotal for both genomic and epigenomic regulations. The regulation of NF-κB by lncRNAs, achieved through both direct and indirect methods affecting NF-κB’s expression or activity, is essential for cellular homeostasis and the response to environmental stimuli. However, dysregulation in this system can lead to the development of various diseases. This review underscores the significant roles of specific lncRNAs in modulating inflammatory responses and macrophage polarization, positioning them as potential biomarkers and therapeutic targets. The divergent roles of certain lncRNAs across different disease contexts underscore the complexity of their functions, thereby emphasizing the need for further research to enhance our understanding and facilitate their clinical application.
Future research should focus on unraveling the detailed mechanisms of how lncRNAs influence NF-κB pathways and macrophage polarization in various diseases. The development of more specific and effective lncRNA-targeted therapies, possibly using advanced techniques like CRISPR/Cas9 and exosome-mediated delivery, is another crucial direction. Clinical trials focusing on the therapeutic potential of lncRNAs and their safety and efficacy in human diseases will be vital. Additionally, exploring the role of lncRNAs in other immune cells and their interaction with macrophages could provide a more comprehensive understanding of immune regulation and disease pathogenesis.

Author Contributions

Conceptualization, W.-H.L. and K.S.; writing—original draft preparation, J.-J.S., H.-S.S., J.P., and I.A.; writing—review and editing, W.-H.L. and K.S.; supervision, W.-H.L.; project administration, W.-H.L.; funding acquisition, W.-H.L. and K.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korean government (Ministry of Science and ICT) (No. 2022R1A2C1010005).

Data Availability Statement

The data used in this article are sourced from materials mentioned in the References section.

Acknowledgments

The authors greatly appreciates the help of Generative AI and AI-Assisted Technologies in the Writing Process. During the preparation of this work, the authors used ChatGPT to improve the quality of English writing. After using this service, the authors reviewed and edited the content as needed and take full responsibility for the content of the publication.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

ABCA1, ATP-binding cassette subfamily A member 1; AML, acute myeloid leukemia; ASncmtRNA, antisense noncoding mitochondrial RNA; ASO, antisense oligonucleotide; BMDM, bone marrow-derived macrophage; BST2, bone marrow stromal cell antigen 2; BTK, Bruton’s tyrosine kinase; CAD, coronary artery disease; Cas, CRISPR-associated protein; COX, cyclooxygenase; CRISPR, clustered regularly interspaced short palindromic repeats; CRNDE, colorectal neoplasia differentially expressed; DDA1, DET1- and DDB1-associated 1; EAE, experimental autoimmune encephalomyelitis; EV, extracellular vesicle; EZH2, enhancer of zeste homolog 2; FGD5-AS1, FGD5 antisense RNA 1; GAS5, growth arrest-specific 5; HCC, hepatocellular carcinoma; HMGB1, high mobility group box 1; HOTAIR, HOX transcript antisense RNA; HOTTIP, HOXA transcript at the distal tip; HUVEC, human umbilical vein endothelial cell; IL, interleukin; LDL, low-density lipoprotein; lncRNA, long noncoding RNA; LPS, lipopolysaccharide; MALAT1, metastasis-associated lung adenocarcinoma transcript 1; MAPK, mitogen-activated protein kinase; MI, myocardial infarction; miRNA, microRNA; MMP, matrix metalloproteinase; MyD88, myeloid differentiation primary response 88; NEAT1, nuclear paraspeckle assembly transcript 1; NF-κB, nuclear factor kappa-light-chain-enhancer of activated B cells; NLRP3, NLR family pyrin domain containing 3; oxLDL, oxidized-LDL; PBMC, peripheral blood mononuclear cell; PINT, P53-induced transcript; PRC2, polycomb repressive complex 2; PVT1, plasmacytoma variant translocation 1; ROS, reactive oxygen species; shRNA, short hairpin RNA; siRNA, small interfering RNA; SMAD3, small mothers against decapentaplegic homolog 3; SNHG, small nucleolar RNA host gene; SOCS, suppressor of cytokine signaling; STAT, signal transducer and activator of transcription; TAK1, TGF-β-activated kinase 1; TAM, tumor-associated macrophage; TLR, Toll-like receptor; TNF-α, tumor necrosis factor-α; TRAF6, TNF receptor-associated factor 6; ULK1, Unc51-like autophagy activating kinase 1; XIST, X-inactive specific transcript.

References

  1. Plytycz, B.; Seljelid, R. From inflammation to sickness: Historical perspective. Arch. Immunol. Ther. Exp. (Warsz) 2003, 51, 105–109. [Google Scholar]
  2. Li, X.; Ma, L. From biological aging to functional decline: Insights into chronic inflammation and intrinsic capacity. Ageing Res. Rev. 2024, 93, 102175. [Google Scholar] [CrossRef]
  3. Lowe, D.B.; Storkus, W.J. Chronic inflammation and immunologic-based constraints in malignant disease. Immunotherapy 2011, 3, 1265–1274. [Google Scholar] [CrossRef]
  4. Furman, D.; Campisi, J.; Verdin, E.; Carrera-Bastos, P.; Targ, S.; Franceschi, C.; Ferrucci, L.; Gilroy, D.W.; Fasano, A.; Miller, G.W.; et al. Chronic inflammation in the etiology of disease across the life span. Nat. Med. 2019, 25, 1822–1832. [Google Scholar] [CrossRef] [PubMed]
  5. Tardito, S.; Martinelli, G.; Soldano, S.; Paolino, S.; Pacini, G.; Patane, M.; Alessandri, E.; Smith, V.; Cutolo, M. Macrophage M1/M2 polarization and rheumatoid arthritis: A systematic review. Autoimmun. Rev. 2019, 18, 102397. [Google Scholar] [CrossRef]
  6. Smolen, J.S.; Feist, E.; Fatenejad, S.; Grishin, S.A.; Korneva, E.V.; Nasonov, E.L.; Samsonov, M.Y.; Fleischmann, R.M.; Group, C. Olokizumab versus Placebo or Adalimumab in Rheumatoid Arthritis. N. Engl. J. Med. 2022, 387, 715–726. [Google Scholar] [CrossRef]
  7. Motwani, M.P.; Gilroy, D.W. Macrophage development and polarization in chronic inflammation. Semin. Immunol. 2015, 27, 257–266. [Google Scholar] [CrossRef]
  8. Bashir, S.; Sharma, Y.; Elahi, A.; Khan, F. Macrophage polarization: The link between inflammation and related diseases. Inflamm. Res. 2016, 65, 1–11. [Google Scholar] [CrossRef] [PubMed]
  9. Jiang, P.; Li, X. Regulatory Mechanism of lncRNAs in M1/M2 Macrophages Polarization in the Diseases of Different Etiology. Front. Immunol. 2022, 13, 835932. [Google Scholar] [CrossRef] [PubMed]
  10. Orecchioni, M.; Ghosheh, Y.; Pramod, A.B.; Ley, K. Macrophage Polarization: Different Gene Signatures in M1(LPS+) vs. Classically and M2(LPS-) vs. Alternatively Activated Macrophages. Front. Immunol. 2019, 10, 1084. [Google Scholar] [CrossRef]
  11. Yunna, C.; Mengru, H.; Lei, W.; Weidong, C. Macrophage M1/M2 polarization. Eur. J. Pharmacol. 2020, 877, 173090. [Google Scholar] [CrossRef] [PubMed]
  12. Funes, S.C.; Rios, M.; Escobar-Vera, J.; Kalergis, A.M. Implications of macrophage polarization in autoimmunity. Immunology 2018, 154, 186–195. [Google Scholar] [CrossRef]
  13. Feng, H.; Zhang, D.; Yin, Y.; Kang, J.; Zheng, R. Salidroside ameliorated the pulmonary inflammation induced by cigarette smoke via mitigating M1 macrophage polarization by JNK/c-Jun. Phytother. Res. 2023, 37, 4251–4264. [Google Scholar] [CrossRef] [PubMed]
  14. Zhao, W.; Ma, L.; Deng, D.; Zhang, T.; Han, L.; Xu, F.; Huang, S.; Ding, Y.; Chen, X. M2 macrophage polarization: A potential target in pain relief. Front. Immunol. 2023, 14, 1243149. [Google Scholar] [CrossRef] [PubMed]
  15. Komal, S.; Han, S.N.; Cui, L.G.; Zhai, M.M.; Zhou, Y.J.; Wang, P.; Shakeel, M.; Zhang, L.R. Epigenetic Regulation of Macrophage Polarization in Cardiovascular Diseases. Pharmaceuticals 2023, 16, 141. [Google Scholar] [CrossRef]
  16. Lawrence, T.; Gilroy, D.W.; Colville-Nash, P.R.; Willoughby, D.A. Possible new role for NF-kappaB in the resolution of inflammation. Nat. Med. 2001, 7, 1291–1297. [Google Scholar] [CrossRef] [PubMed]
  17. Napetschnig, J.; Wu, H. Molecular basis of NF-kappaB signaling. Annu. Rev. Biophys. 2013, 42, 443–468. [Google Scholar] [CrossRef]
  18. Yu, H.; Lin, L.; Zhang, Z.; Zhang, H.; Hu, H. Targeting NF-kappaB pathway for the therapy of diseases: Mechanism and clinical study. Signal Transduct. Target. Ther. 2020, 5, 209. [Google Scholar] [CrossRef]
  19. Khorkova, O.; Hsiao, J.; Wahlestedt, C. Basic biology and therapeutic implications of lncRNA. Adv. Drug Deliv. Rev. 2015, 87, 15–24. [Google Scholar] [CrossRef]
  20. Herman, A.B.; Tsitsipatis, D.; Gorospe, M. Integrated lncRNA function upon genomic and epigenomic regulation. Mol. Cell 2022, 82, 2252–2266. [Google Scholar] [CrossRef]
  21. Ferrè, F.; Colantoni, A.; Helmer-Citterich, M. Revealing protein-lncRNA interaction. Brief. Bioinform. 2016, 17, 106–116. [Google Scholar] [CrossRef] [PubMed]
  22. Peng, W.X.; Koirala, P.; Mo, Y.Y. LncRNA-mediated regulation of cell signaling in cancer. Oncogene 2017, 36, 5661–5667. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, C.; Wang, L.; Ding, Y.; Lu, X.; Zhang, G.; Yang, J.; Zheng, H.; Wang, H.; Jiang, Y.; Xu, L. LncRNA Structural Characteristics in Epigenetic Regulation. Int. J. Mol. Sci. 2017, 18, 2659. [Google Scholar] [CrossRef] [PubMed]
  24. Senmatsu, S.; Hirota, K. Roles of lncRNA transcription as a novel regulator of chromosomal function. Genes. Genet. Syst. 2021, 95, 213–223. [Google Scholar] [CrossRef] [PubMed]
  25. Huang, W.; Li, H.; Yu, Q.; Xiao, W.; Wang, D.O. LncRNA-mediated DNA methylation: An emerging mechanism in cancer and beyond. J. Exp. Clin. Cancer Res. 2022, 41, 100. [Google Scholar] [CrossRef]
  26. Mathieu, E.L.; Belhocine, M.; Dao, L.T.; Puthier, D.; Spicuglia, S. [Functions of lncRNA in development and diseases]. Med. Sci. 2014, 30, 790–796. [Google Scholar] [CrossRef]
  27. Zhao, Z.; Sun, W.; Guo, Z.; Zhang, J.; Yu, H.; Liu, B. Mechanisms of lncRNA/microRNA interactions in angiogenesis. Life Sci. 2020, 254, 116900. [Google Scholar] [CrossRef]
  28. Pravda, J. Sepsis: Evidence-based pathogenesis and treatment. World J. Crit. Care Med. 2021, 10, 66–80. [Google Scholar] [CrossRef]
  29. Cristofaro, P.; Opal, S.M. The Toll-like receptors and their role in septic shock. Expert. Opin. Ther. Targets 2003, 7, 603–612. [Google Scholar] [CrossRef]
  30. Chousterman, B.G.; Swirski, F.K.; Weber, G.F. Cytokine storm and sepsis disease pathogenesis. Semin. Immunopathol. 2017, 39, 517–528. [Google Scholar] [CrossRef] [PubMed]
  31. Tracey, K.J.; Beutler, B.; Lowry, S.F.; Merryweather, J.; Wolpe, S.; Milsark, I.W.; Hariri, R.J.; Fahey, T.J., 3rd; Zentella, A.; Albert, J.D.; et al. Shock and tissue injury induced by recombinant human cachectin. Science 1986, 234, 470–474. [Google Scholar] [CrossRef] [PubMed]
  32. Chen, X.; Liu, Y.; Gao, Y.; Shou, S.; Chai, Y. The roles of macrophage polarization in the host immune response to sepsis. Int. Immunopharmacol. 2021, 96, 107791. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, S.F.; Malik, A.B. NF-kappa B activation as a pathological mechanism of septic shock and inflammation. Am. J. Physiol. Lung Cell Mol. Physiol. 2006, 290, L622–L645. [Google Scholar] [CrossRef]
  34. Naganuma, T.; Nakagawa, S.; Tanigawa, A.; Sasaki, Y.F.; Goshima, N.; Hirose, T. Alternative 3’-end processing of long noncoding RNA initiates construction of nuclear paraspeckles. EMBO J. 2012, 31, 4020–4034. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, C.C.; Niu, F. LncRNA NEAT1 promotes inflammatory response in sepsis-induced liver injury via the Let-7a/TLR4 axis. Int. Immunopharmacol. 2019, 75, 105731. [Google Scholar] [CrossRef] [PubMed]
  36. Xia, D.; Yao, R.; Zhou, P.; Wang, C.; Xia, Y.; Xu, S. LncRNA NEAT1 reversed the hindering effects of miR-495-3p/STAT3 axis and miR-211/PI3K/AKT axis on sepsis-relevant inflammation. Mol. Immunol. 2020, 117, 168–179. [Google Scholar] [CrossRef] [PubMed]
  37. Wu, X.Y.; Fang, Y.; Zheng, F.X.; Zhang, Y.Z.; Li, Q.L. LncRNA NEAT1 facilitates the progression of sepsis through up-regulating TSP-1 via sponging miR-370-3p. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 333–344. [Google Scholar] [CrossRef]
  38. Li, Y.; Guo, W.; Cai, Y. NEAT1 Promotes LPS-induced Inflammatory Injury in Macrophages by Regulating MiR-17-5p/TLR4. Open Med. 2020, 15, 38–49. [Google Scholar] [CrossRef]
  39. Yang, Y.; Xue, J.; Qin, L.; Zhang, J.; Liu, J.; Yu, J. LncRNA NEAT1 Promotes Inflammatory Response in Sepsis via the miR-31-5p/POU2F1 Axis. Inflammation 2021, 44, 1518–1528. [Google Scholar] [CrossRef]
  40. Wang, W.; Guo, Z.H. Downregulation of lncRNA NEAT1 Ameliorates LPS-Induced Inflammatory Responses by Promoting Macrophage M2 Polarization via miR-125a-5p/TRAF6/TAK1 Axis. Inflammation 2020, 43, 1548–1560. [Google Scholar] [CrossRef]
  41. Chen, J.; Tang, S.; Ke, S.; Cai, J.J.; Osorio, D.; Golovko, A.; Morpurgo, B.; Guo, S.; Sun, Y.; Winkle, M.; et al. Ablation of long noncoding RNA MALAT1 activates antioxidant pathway and alleviates sepsis in mice. Redox Biol. 2022, 54, 102377. [Google Scholar] [CrossRef]
  42. Zhao, G.; Su, Z.; Song, D.; Mao, Y.; Mao, X. The long noncoding RNA MALAT1 regulates the lipopolysaccharide-induced inflammatory response through its interaction with NF-kappaB. FEBS Lett. 2016, 590, 2884–2895. [Google Scholar] [CrossRef] [PubMed]
  43. Dai, L.; Zhang, G.; Cheng, Z.; Wang, X.; Jia, L.; Jing, X.; Wang, H.; Zhang, R.; Liu, M.; Jiang, T.; et al. Knockdown of LncRNA MALAT1 contributes to the suppression of inflammatory responses by up-regulating miR-146a in LPS-induced acute lung injury. Connect. Tissue Res. 2018, 59, 581–592. [Google Scholar] [CrossRef]
  44. Lin, L.P.; Niu, G.H.; Zhang, X.Q. Influence of lncRNA MALAT1 on septic lung injury in mice through p38 MAPK/p65 NF-kappaB pathway. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 1296–1304. [Google Scholar] [CrossRef] [PubMed]
  45. Cui, H.; Banerjee, S.; Guo, S.; Xie, N.; Ge, J.; Jiang, D.; Zornig, M.; Thannickal, V.J.; Liu, G. Long noncoding RNA Malat1 regulates differential activation of macrophages and response to lung injury. JCI Insight. 2019, 4, 124522. [Google Scholar] [CrossRef] [PubMed]
  46. Yang, Q.; Cao, K.; Jin, G.; Zhang, J. Hsa-miR-346 plays a role in the development of sepsis by downregulating SMAD3 expression and is negatively regulated by lncRNA MALAT1. Mol. Cell Probes 2019, 47, 101444. [Google Scholar] [CrossRef] [PubMed]
  47. Chen, J.; Ren, H.; Liu, B. Evaluating the potency of blood long noncoding RNA PVT1 as candidate biomarker reflecting inflammation, multiple organ dysfunction, and mortality risk in sepsis patients. J. Clin. Lab. Anal. 2022, 36, e24268. [Google Scholar] [CrossRef]
  48. Yuan, W.; Xiong, X.; Du, J.; Fan, Q.; Wang, R.; Zhang, X. LncRNA PVT1 accelerates LPS-induced septic acute kidney injury through targeting miR-17-5p and regulating NF-kappaB pathway. Int. Urol. Nephrol. 2021, 53, 2409–2419. [Google Scholar] [CrossRef]
  49. Zheng, S.; Li, W.; Liao, W.; Huang, C.; Zhou, M.; Zheng, Y.; Zou, Z.; He, Z. Silencing of LncRNA-PVT1 ameliorates lipopolysaccharide-induced inflammation in THP-1-derived macrophages via inhibition of the p38 MAPK signaling pathway. Ann. Palliat. Med. 2021, 10, 6410–6418. [Google Scholar] [CrossRef]
  50. Luo, Y.Y.; Yang, Z.Q.; Lin, X.F.; Zhao, F.L.; Tu, H.T.; Wang, L.J.; Wen, M.Y.; Xian, S.X. Knockdown of lncRNA PVT1 attenuated macrophage M1 polarization and relieved sepsis induced myocardial injury via miR-29a/HMGB1 axis. Cytokine 2021, 143, 155509. [Google Scholar] [CrossRef]
  51. Hreggvidsdottir, H.S.; Lundberg, A.M.; Aveberger, A.C.; Klevenvall, L.; Andersson, U.; Harris, H.E. High mobility group box protein 1 (HMGB1)-partner molecule complexes enhance cytokine production by signaling through the partner molecule receptor. Mol. Med. 2012, 18, 224–230. [Google Scholar] [CrossRef]
  52. Meng, Y.; Qiu, S.; Sun, L.; Zuo, J. Knockdown of exosome-mediated lnc-PVT1 alleviates lipopolysaccharide-induced osteoarthritis progression by mediating the HMGB1/TLR4/NF-kappaB pathway via miR-93-5p. Mol. Med. Rep. 2020, 22, 5313–5325. [Google Scholar] [CrossRef]
  53. Pan, X.; He, L. LncRNA MEG3 expression in sepsis and its effect on LPS-induced macrophage function. Cell. Mol. Biol. (Noisy-le-grand) 2020, 66, 131–136. [Google Scholar] [CrossRef]
  54. Wang, Y.; Xu, Z.; Yue, D.; Zeng, Z.; Yuan, W.; Xu, K. Linkage of lncRNA CRNDE sponging miR-181a-5p with aggravated inflammation underlying sepsis. Innate Immun. 2020, 26, 152–161. [Google Scholar] [CrossRef]
  55. Gibson, M.S.; Domingues, N.; Vieira, O.V. Lipid and Non-lipid Factors Affecting Macrophage Dysfunction and Inflammation in Atherosclerosis. Front. Physiol. 2018, 9, 654. [Google Scholar] [CrossRef] [PubMed]
  56. Lusis, A.J. Atherosclerosis. Nature 2000, 14, 233–241. [Google Scholar] [CrossRef]
  57. Libby, P. Atherosclerosis in Inflammation. Nature 2002, 420, 868–874. [Google Scholar] [CrossRef] [PubMed]
  58. Ross, R. Atherosclerosis--an inflammatory disease. N. Engl. J. Med. 1999, 340, 115–126. [Google Scholar] [CrossRef]
  59. Wu, J.; He, S.; Song, Z.; Chen, S.; Lin, X.; Sun, H.; Zhou, P.; Peng, Q.; Du, S.; Zheng, S.; et al. Macrophage polarization states in atherosclerosis. Front. Immunol. 2023, 14, 1185587. [Google Scholar] [CrossRef]
  60. Momtazi-Borojeni, A.A.; Abdollahi, E.; Nikfar, B.; Chaichian, S.; Ekhlasi-Hundrieser, M. Curcumin as a potential modulator of M1 and M2 macrophages: New insights in atherosclerosis therapy. Heart Fail. Rev. 2019, 24, 399–409. [Google Scholar] [CrossRef] [PubMed]
  61. Park, S.H. Regulation of Macrophage Activation and Differentiation in Atherosclerosis. J. Lipid Atheroscler. 2021, 10, 251–267. [Google Scholar] [CrossRef]
  62. Tabas, I. Macrophage death and defective inflammation resolution in atherosclerosis. Nat. Rev. Immunol. 2010, 10, 36–46. [Google Scholar] [CrossRef]
  63. Jiang, H.; Zhou, Y.; Nabavi, S.M.; Sahebkar, A.; Little, P.J.; Xu, S.; Weng, J.; Ge, J. Mechanisms of Oxidized LDL-Mediated Endothelial Dysfunction and Its Consequences for the Development of Atherosclerosis. Front. Cardiovasc. Med. 2022, 9, 925923. [Google Scholar] [CrossRef]
  64. Watanabe, N.; Ikeda, U. Matrix metalloproteinases and atherosclerosis. Curr. Atheroscler. Rep. 2004, 6, 112–120. [Google Scholar] [CrossRef]
  65. Wilson, H.M. The intracellular signaling pathways governing macrophage activation and function in human atherosclerosis. Biochem. Soc. Trans. 2022, 50, 1673–1682. [Google Scholar] [CrossRef]
  66. Taghizadeh, E.; Taheri, F.; Renani, P.G.; Reiner, Z.; Navashenaq, J.G.; Sahebkar, A. Macrophage: A Key Therapeutic Target in Atherosclerosis? Curr. Pharm. Des. 2019, 25, 3165–3174. [Google Scholar] [CrossRef]
  67. Vlachogiannis, N.I.; Sachse, M.; Georgiopoulos, G.; Zormpas, E.; Bampatsias, D.; Delialis, D.; Bonini, F.; Galyfos, G.; Sigala, F.; Stamatelopoulos, K.; et al. Adenosine-to-inosine Alu RNA editing controls the stability of the pro-inflammatory long noncoding RNA NEAT1 in atherosclerotic cardiovascular disease. J. Mol. Cell Cardiol. 2021, 160, 111–120. [Google Scholar] [CrossRef] [PubMed]
  68. Guo, J.T.; Wang, L.; Yu, H.B. Knockdown of NEAT1 mitigates ox-LDL-induced injury in human umbilical vein endothelial cells via miR-30c-5p/TCF7 axis. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 9633–9644. [Google Scholar] [CrossRef] [PubMed]
  69. Huang-Fu, N.; Cheng, J.S.; Wang, Y.; Li, Z.W.; Wang, S.H. Neat1 regulates oxidized low-density lipoprotein-induced inflammation and lipid uptake in macrophages via paraspeckle formation. Mol. Med. Rep. 2018, 17, 3092–3098. [Google Scholar] [CrossRef] [PubMed]
  70. Wang, L.; Xia, J.W.; Ke, Z.P.; Zhang, B.H. Blockade of NEAT1 represses inflammation response and lipid uptake via modulating miR-342-3p in human macrophages THP-1 cells. J. Cell Physiol. 2019, 234, 5319–5326. [Google Scholar] [CrossRef] [PubMed]
  71. Lin, S.; Wen, Z.; Li, S.; Chen, Z.; Li, C.; Ouyang, Z.; Lin, C.; Kuang, M.; Xue, C.; Ding, Y. LncRNA Neat1 promotes the macrophage inflammatory response and acts as a therapeutic target in titanium particle-induced osteolysis. Acta Biomater. 2022, 142, 345–360. [Google Scholar] [CrossRef] [PubMed]
  72. Chen, D.D.; Hui, L.L.; Zhang, X.C.; Chang, Q. NEAT1 contributes to ox-LDL-induced inflammation and oxidative stress in macrophages through inhibiting miR-128. J. Cell Biochem. 2018, 120, 2493–2501. [Google Scholar] [CrossRef] [PubMed]
  73. Gast, M.; Rauch, B.H.; Haghikia, A.; Nakagawa, S.; Haas, J.; Stroux, A.; Schmidt, D.; Schumann, P.; Weiss, S.; Jensen, L.; et al. Long noncoding RNA NEAT1 modulates immune cell functions and is suppressed in early onset myocardial infarction patients. Cardiovasc. Res. 2019, 115, 1886–1906. [Google Scholar] [CrossRef] [PubMed]
  74. Du, H.; Zhang, H.; Yang, R.; Qiao, L.; Shao, H.; Zhang, X. Small interfering RNA-induced silencing lncRNA PVT1 inhibits atherosclerosis via inactivating the MAPK/NF-kappaB pathway. Aging 2021, 13, 24449–24463. [Google Scholar] [CrossRef] [PubMed]
  75. An, J.H.; Chen, Z.Y.; Ma, Q.L.; Wang, H.J.; Zhang, J.Q.; Shi, F.W. LncRNA SNHG16 promoted proliferation and inflammatory response of macrophages through miR-17-5p/NF-kappaB signaling pathway in patients with atherosclerosis. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 8665–8677. [Google Scholar] [CrossRef] [PubMed]
  76. Yang, K.; Xue, Y.; Gao, X. LncRNA XIST Promotes Atherosclerosis by Regulating miR-599/TLR4 Axis. Inflammation 2021, 44, 965–973. [Google Scholar] [CrossRef]
  77. Sun, C.; Fu, Y.; Gu, X.; Xi, X.; Peng, X.; Wang, C.; Sun, Q.; Wang, X.; Qian, F.; Qin, Z.; et al. Macrophage-Enriched lncRNA RAPIA: A Novel Therapeutic Target for Atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2020, 40, 1464–1478. [Google Scholar] [CrossRef]
  78. Gareev, I.; Kudriashov, V.; Sufianov, A.; Begliarzade, S.; Ilyasova, T.; Liang, Y.; Beylerli, O. The role of long non-coding RNA ANRIL in the development of atherosclerosis. Noncoding RNA Res. 2022, 7, 212–216. [Google Scholar] [CrossRef]
  79. Zhang, L.; Cheng, H.; Yue, Y.; Li, S.; Zhang, D.; He, R. H19 knockdown suppresses proliferation and induces apoptosis by regulating miR-148b/WNT/beta-catenin in ox-LDL -stimulated vascular smooth muscle cells. J. Biomed. Sci. 2018, 25, 11. [Google Scholar] [CrossRef]
  80. Lu, G.; Chu, Y.; Tian, P. Knockdown of H19 Attenuates Ox-LDL-induced Vascular Smooth Muscle Cell Proliferation, Migration, and Invasion by Regulating miR-599/PAPPA Axis. J. Cardiovasc. Pharmacol. 2021, 77, 386–396. [Google Scholar] [CrossRef]
  81. Safaei, S.; Tahmasebi-Birgani, M.; Bijanzadeh, M.; Seyedian, S.M. Increased Expression Level of Long Noncoding RNA H19 in Plasma of Patients with Myocardial Infarction. Int. J. Mol. Cell Med. 2020, 9, 122–129. [Google Scholar] [CrossRef]
  82. Zhang, Z.; Gao, W.; Long, Q.Q.; Zhang, J.; Li, Y.F.; Liu, D.C.; Yan, J.J.; Yang, Z.J.; Wang, L.S. Increased plasma levels of lncRNA H19 and LIPCAR are associated with increased risk of coronary artery disease in a Chinese population. Sci. Rep. 2017, 7, 7491. [Google Scholar] [CrossRef] [PubMed]
  83. Han, Y.; Ma, J.; Wang, J.; Wang, L. Silencing of H19 inhibits the adipogenesis and inflammation response in ox-LDL-treated Raw264.7 cells by up-regulating miR-130b. Mol. Immunol. 2018, 93, 107–114. [Google Scholar] [CrossRef] [PubMed]
  84. Cao, L.; Zhang, Z.; Li, Y.; Zhao, P.; Chen, Y. LncRNA H19/miR-let-7 axis participates in the regulation of ox-LDL-induced endothelial cell injury via targeting periostin. Int. Immunopharmacol. 2019, 72, 496–503. [Google Scholar] [CrossRef]
  85. Wang, X.; Che, Y.; Wang, J.; Men, K. [Effects and mechanism of knocking down lncRNA H19 to inhibit lipid accumulation in human THP-1 cells-derived macrophages]. Xi Bao Yu Fen. Zi Mian Yi Xue Za Zhi 2023, 39, 884–890. [Google Scholar] [PubMed]
  86. Pan, J.X. LncRNA H19 promotes atherosclerosis by regulating MAPK and NF-kB signaling pathway. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 322–328. [Google Scholar] [PubMed]
  87. Qiu, S.; Sun, J. lncRNA-MALAT1 expression in patients with coronary atherosclerosis and its predictive value for in-stent restenosis. Exp. Ther. Med. 2020, 20, 129. [Google Scholar] [CrossRef]
  88. Liu, L.; Tan, L.; Yao, J.; Yang, L. Long non-coding RNA MALAT1 regulates cholesterol accumulation in ox-LDL-induced macrophages via the microRNA-17-5p/ABCA1 axis. Mol. Med. Rep. 2020, 21, 1761–1770. [Google Scholar] [CrossRef]
  89. Wang, L.; Qi, Y.; Wang, Y.; Tang, H.; Li, Z.; Wang, Y.; Tang, S.; Zhu, H. LncRNA MALAT1 Suppression Protects Endothelium against oxLDL-Induced Inflammation via Inhibiting Expression of MiR-181b Target Gene TOX. Oxid. Med. Cell Longev. 2019, 2019, 8245810. [Google Scholar] [CrossRef]
  90. Huangfu, N.; Xu, Z.; Zheng, W.; Wang, Y.; Cheng, J.; Chen, X. LncRNA MALAT1 regulates oxLDL-induced CD36 expression via activating beta-catenin. Biochem. Biophys. Res. Commun. 2018, 495, 2111–2117. [Google Scholar] [CrossRef]
  91. Shi, Z.; Zheng, Z.; Lin, X.; Ma, H. Long Noncoding RNA MALAT1 Regulates the Progression of Atherosclerosis by miR-330-5p/NF-kappaB Signal Pathway. J. Cardiovasc. Pharmacol. 2021, 78, 235–246. [Google Scholar] [CrossRef]
  92. Ma, Z.; Zhang, J.; Xu, X.; Qu, Y.; Dong, H.; Dang, J.; Huo, Z.; Xu, G. LncRNA expression profile during autophagy and Malat1 function in macrophages. PLoS ONE 2019, 14, e0221104. [Google Scholar] [CrossRef]
  93. Yang, J.; Lin, X.; Wang, L.; Sun, T.; Zhao, Q.; Ma, Q.; Zhou, Y. LncRNA MALAT1 Enhances ox-LDL-Induced Autophagy through the SIRT1/MAPK/NF-kappaB Pathway in Macrophages. Curr. Vasc. Pharmacol. 2020, 18, 652–662. [Google Scholar] [CrossRef] [PubMed]
  94. Rusetskaya, N.Y.; Loginova, N.Y.; Pokrovskaya, E.P.; Chesovskikh, Y.S.; Titova, L.E. Redox regulation of the NLRP3-mediated inflammation and pyroptosis. Biomed. Khim 2023, 69, 333–352. [Google Scholar] [CrossRef] [PubMed]
  95. Han, Y.; Qiu, H.; Pei, X.; Fan, Y.; Tian, H.; Geng, J. Low-dose Sinapic Acid Abates the Pyroptosis of Macrophages by Downregulation of lncRNA-MALAT1 in Rats With Diabetic Atherosclerosis. J. Cardiovasc. Pharmacol. 2018, 71, 104–112. [Google Scholar] [CrossRef] [PubMed]
  96. Zheng, T. A Review of the Roles of Specialized Extracellular Vesicles, Migrasomes, and Exosomes in Normal Cell Physiology and Disease. Med. Sci. Monit. 2023, 29, e940118. [Google Scholar] [CrossRef] [PubMed]
  97. Chen, B.; Luo, L.; Wei, X.; Gong, D.; Li, Z.; Li, S.; Tang, W.; Jin, L. M1 Bone Marrow-Derived Macrophage-Derived Extracellular Vesicles Inhibit Angiogenesis and Myocardial Regeneration Following Myocardial Infarction via the MALAT1/MicroRNA-25-3p/CDC42 Axis. Oxidative Med. Cell. Longev. 2021, 2021, 9959746. [Google Scholar] [CrossRef] [PubMed]
  98. Afonso Mda, S.; Castilho, G.; Lavrador, M.S.; Passarelli, M.; Nakandakare, E.R.; Lottenberg, S.A.; Lottenberg, A.M. The impact of dietary fatty acids on macrophage cholesterol homeostasis. J. Nutr. Biochem. 2014, 25, 95–103. [Google Scholar] [CrossRef] [PubMed]
  99. Gast, M.; Rauch, B.H.; Nakagawa, S.; Haghikia, A.; Jasina, A.; Haas, J.; Nath, N.; Jensen, L.; Stroux, A.; Bohm, A.; et al. Immune system-mediated atherosclerosis caused by deficiency of long non-coding RNA MALAT1 in ApoE-/-mice. Cardiovasc. Res. 2019, 115, 302–314. [Google Scholar] [CrossRef]
  100. Huang, C.; Han, J.; Wu, Y.; Li, S.; Wang, Q.; Lin, W.; Zhu, J. Exosomal MALAT1 derived from oxidized low-density lipoprotein-treated endothelial cells promotes M2 macrophage polarization. Mol. Med. Rep. 2018, 18, 509–515. [Google Scholar] [CrossRef]
  101. Pang, J.L.; Wang, J.W.; Hu, P.Y.; Jiang, J.S.; Yu, C. HOTAIR alleviates ox-LDL-induced inflammatory response in Raw264.7 cells via inhibiting NF-kappaB pathway. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 6991–6998. [Google Scholar] [CrossRef]
  102. Tamanini, F.; Bontekoe, C.; Bakker, C.E.; van Unen, L.; Anar, B.; Willemsen, R.; Yoshida, M.; Galjaard, H.; Oostra, B.A.; Hoogeveen, A.T. Different targets for the fragile X-related proteins revealed by their distinct nuclear localizations. Hum. Mol. Genet. 1999, 8, 863–869. [Google Scholar] [CrossRef]
  103. Castellanos-Rubio, A.; Kratchmarov, R.; Sebastian, M.; Garcia-Etxebarria, K.; Garcia, L.; Irastorza, I.; Ghosh, S. Cytoplasmic Form of Carlr lncRNA Facilitates Inflammatory Gene Expression upon NF-kappaB Activation. J. Immunol. 2017, 199, 581–588. [Google Scholar] [CrossRef] [PubMed]
  104. Xue, Z.; Zhang, Z.; Liu, H.; Li, W.; Guo, X.; Zhang, Z.; Liu, Y.; Jia, L.; Li, Y.; Ren, Y.; et al. lincRNA-Cox2 regulates NLRP3 inflammasome and autophagy mediated neuroinflammation. Cell Death Differ. 2019, 26, 130–145. [Google Scholar] [CrossRef] [PubMed]
  105. Hu, G.; Gong, A.Y.; Wang, Y.; Ma, S.; Chen, X.; Chen, J.; Su, C.J.; Shibata, A.; Strauss-Soukup, J.K.; Drescher, K.M.; et al. LincRNA-Cox2 Promotes Late Inflammatory Gene Transcription in Macrophages through Modulating SWI/SNF-Mediated Chromatin Remodeling. J. Immunol. 2016, 196, 2799–2808. [Google Scholar] [CrossRef]
  106. Krawczyk, M.; Emerson, B.M. p50-associated COX-2 extragenic RNA (PACER) activates COX-2 gene expression by occluding repressive NF-kappaB complexes. eLife 2014, 3, e01776. [Google Scholar] [CrossRef] [PubMed]
  107. Ye, M.; Wang, C.; Zhu, J.; Chen, M.; Wang, S.; Li, M.; Lu, Y.; Xiao, P.; Zhou, M.; Li, X.; et al. An NF-kappaB-responsive long noncoding RNA, PINT, regulates TNF-alpha gene transcription by scaffolding p65 and EZH2. FASEB J. 2021, 35, e21667. [Google Scholar] [CrossRef]
  108. Wu, Q.; Han, L.; Yan, W.; Ji, X.; Han, R.; Yang, J.; Yuan, J.; Ni, C. miR-489 inhibits silica-induced pulmonary fibrosis by targeting MyD88 and Smad3 and is negatively regulated by lncRNA CHRF. Sci. Rep. 2016, 6, 30921. [Google Scholar] [CrossRef]
  109. Shen, M.; Pan, X.; Gao, Y.; Ye, H.; Zhang, J.; Chen, Y.; Pan, M.; Huang, W.; Xu, X.; Zhao, Y.; et al. LncRNA CRNDE Exacerbates IgA Nephropathy Progression by Promoting NLRP3 Inflammasome Activation in Macrophages. Immunol. Investig. 2022, 51, 1515–1527. [Google Scholar] [CrossRef]
  110. Wen, S.; Li, F.; Tang, Y.; Dong, L.; He, Y.; Deng, Y.; Tao, Z. MIR222HG attenuates macrophage M2 polarization and allergic inflammation in allergic rhinitis by targeting the miR146a-5p/TRAF6/NF-kappaB axis. Front. Immunol. 2023, 14, 1168920. [Google Scholar] [CrossRef]
  111. Akincilar, S.C.; Wu, L.; Ng, Q.F.; Chua, J.Y.H.; Unal, B.; Noda, T.; Chor, W.H.J.; Ikawa, M.; Tergaonkar, V. NAIL: An evolutionarily conserved lncRNA essential for licensing coordinated activation of p38 and NFkappaB in colitis. Gut 2021, 70, 1857–1871. [Google Scholar] [CrossRef] [PubMed]
  112. Li, Q.; Lu, L.; Li, X.; Lu, S. Long non-coding RNA NKILA alleviates airway inflammation in asthmatic mice by promoting M2 macrophage polarization and inhibiting the NF-kappaB pathway. Biochem. Biophys. Res. Commun. 2021, 571, 46–52. [Google Scholar] [CrossRef] [PubMed]
  113. Yang, T.; Li, S.; Liu, J.; Yin, D.; Yang, X.; Tang, Q. lncRNA-NKILA/NF-kappaB feedback loop modulates laryngeal cancer cell proliferation, invasion, and radioresistance. Cancer Med. 2018, 7, 2048–2063. [Google Scholar] [CrossRef] [PubMed]
  114. Hu, C.; Li, J.; Tan, Y.; Liu, Y.; Bai, C.; Gao, J.; Zhao, S.; Yao, M.; Lu, X.; Qiu, L.; et al. Tanreqing Injection Attenuates Macrophage Activation and the Inflammatory Response via the lncRNA-SNHG1/HMGB1 Axis in Lipopolysaccharide-Induced Acute Lung Injury. Front. Immunol. 2022, 13, 820718. [Google Scholar] [CrossRef]
  115. Ai, Y.; Liu, S.; Luo, H.; Wu, S.; Wei, H.; Tang, Z.; Li, X.; Zou, C. lncRNA DCST1-AS1 facilitates oral squamous cell carcinoma by promoting M2 macrophage polarization through activating NF-κB signaling. J. Immunol. Res. 2021, 2021, 5524231. [Google Scholar] [CrossRef]
  116. Liu, X.; Li, C.; Zhu, J.; Li, W.; Zhu, Q. Dysregulation of FTX/miR-545 signaling pathway downregulates Tim-3 and is responsible for the abnormal activation of macrophage in cirrhosis. J. Cell. Biochem. 2019, 120, 2336–2346. [Google Scholar] [CrossRef]
  117. Obaid, M.; Udden, S.M.N.; Alluri, P.; Mandal, S.S. LncRNA HOTAIR regulates glucose transporter Glut1 expression and glucose uptake in macrophages during inflammation. Sci. Rep. 2021, 11, 232. [Google Scholar] [CrossRef]
  118. Hu, Y.L.; Feng, Y.; Chen, Y.Y.; Liu, J.Z.; Su, Y.; Li, P.; Huang, H.; Mao, Q.S.; Xue, W.J. SNHG16/miR-605-3p/TRAF6/NF-kappaB feedback loop regulates hepatocellular carcinoma metastasis. J. Cell Mol. Med. 2020, 24, 7637–7651. [Google Scholar] [CrossRef]
  119. Huang, L.; Xiong, S.; Liu, H.; Zhang, R.; Wu, Y.; Hu, X. Silencing LncRNA SNHG16 suppresses the diabetic inflammatory response by targeting the miR-212-3p/NF-kappaB signaling pathway. Diabetol. Metab. Syndr. 2023, 15, 119. [Google Scholar] [CrossRef]
  120. He, Z.; Wang, J.; Zhu, C.; Xu, J.; Chen, P.; Jiang, X.; Chen, Y.; Jiang, J.; Sun, C. Exosome-derived FGD5-AS1 promotes tumor-associated macrophage M2 polarization-mediated pancreatic cancer cell proliferation and metastasis. Cancer Lett. 2022, 548, 215751. [Google Scholar] [CrossRef] [PubMed]
  121. Shen, T.; Miao, S.; Zhou, Y.; Yi, X.; Xue, S.; Du, B.; Tang, C.; Qu, L.; Fu, D.; Jia, R.; et al. Exosomal AP000439.2 from clear cell renal cell carcinoma induces M2 macrophage polarization to promote tumor progression through activation of STAT3. Cell Commun. Signal 2022, 20, 152. [Google Scholar] [CrossRef]
  122. Zhu, X.; Sun, Y.; Yu, Q.; Wang, X.; Wang, Y.; Zhao, Y. Exosomal lncRNA GAS5 promotes M1 macrophage polarization in allergic rhinitis via restraining mTORC1/ULK1/ATG13-mediated autophagy and subsequently activating NF-small ka, CyrillicB signaling. Int. Immunopharmacol. 2023, 121, 110450. [Google Scholar] [CrossRef]
  123. Jiang, H.; Zhou, L.; Shen, N.; Ning, X.; Wu, D.; Jiang, K.; Huang, X. M1 macrophage-derived exosomes and their key molecule lncRNA HOTTIP suppress head and neck squamous cell carcinoma progression by upregulating the TLR5/NF-kappaB pathway. Cell Death Dis. 2022, 13, 183. [Google Scholar] [CrossRef]
  124. Liu, J.; Niu, Z.; Zhang, R.; Peng, Z.; Wang, L.; Liu, Z.; Gao, Y.; Pei, H.; Pan, L. MALAT1 shuttled by extracellular vesicles promotes M1 polarization of macrophages to induce acute pancreatitis via miR-181a-5p/HMGB1 axis. J. Cell Mol. Med. 2021, 25, 9241–9254. [Google Scholar] [CrossRef]
  125. Liu, J.; Geng, X.; Hou, J.; Wu, G. New insights into M1/M2 macrophages: Key modulators in cancer progression. Cancer Cell Int. 2021, 21, 389. [Google Scholar] [CrossRef] [PubMed]
  126. Italiani, P.; Boraschi, D. From Monocytes to M1/M2 Macrophages: Phenotypical vs. Functional Differentiation. Front. Immunol. 2014, 5, 514. [Google Scholar] [CrossRef] [PubMed]
  127. Solinas, G.; Germano, G.; Mantovani, A.; Allavena, P. Tumor-associated macrophages (TAM) as major players of the cancer-related inflammation. J. Leukoc. Biol. 2009, 86, 1065–1073. [Google Scholar] [CrossRef] [PubMed]
  128. Erreni, M.; Mantovani, A.; Allavena, P. Tumor-associated Macrophages (TAM) and Inflammation in Colorectal Cancer. Cancer Microenviron. 2011, 4, 141–154. [Google Scholar] [CrossRef] [PubMed]
  129. Karin, M.; Greten, F.R. NF-kappaB: Linking inflammation and immunity to cancer development and progression. Nat. Rev. Immunol. 2005, 5, 749–759. [Google Scholar] [CrossRef] [PubMed]
  130. Wang, B.; Cheng, D.; Ma, D.; Chen, R.; Li, D.; Zhao, W.; Fang, C.; Ji, M. Mutual regulation of PD-L1 immunosuppression between tumor-associated macrophages and tumor cells: A critical role for exosomes. Cell Commun. Signal 2024, 22, 21. [Google Scholar] [CrossRef] [PubMed]
  131. Seo, K.H.; Ko, H.M.; Choi, J.H.; Jung, H.H.; Chun, Y.H.; Choi, I.W.; Lee, H.K.; Im, S.Y. Essential role for platelet-activating factor-induced NF-kappaB activation in macrophage-derived angiogenesis. Eur. J. Immunol. 2004, 34, 2129–2137. [Google Scholar] [CrossRef]
  132. Mancino, A.; Lawrence, T. Nuclear factor-kappaB and tumor-associated macrophages. Clin. Cancer Res. 2010, 16, 784–789. [Google Scholar] [CrossRef] [PubMed]
  133. Varney, M.L.; Olsen, K.J.; Mosley, R.L.; Bucana, C.D.; Talmadge, J.E.; Singh, R.K. Monocyte/macrophage recruitment, activation and differentiation modulate interleukin-8 production: A paracrine role of tumor-associated macrophages in tumor angiogenesis. In Vivo 2002, 16, 471–477. [Google Scholar] [PubMed]
  134. Hoesel, B.; Schmid, J.A. The complexity of NF-kappaB signaling in inflammation and cancer. Mol. Cancer 2013, 12, 86. [Google Scholar] [CrossRef] [PubMed]
  135. Shang, S.; Ji, X.; Zhang, L.; Chen, J.; Li, C.; Shi, R.; Xiang, W.; Kang, X.; Zhang, D.; Yang, F.; et al. Macrophage ABHD5 Suppresses NFkappaB-Dependent Matrix Metalloproteinase Expression and Cancer Metastasis. Cancer Res. 2019, 79, 5513–5526. [Google Scholar] [CrossRef]
  136. Chen, J.; Wu, D.; Zhang, Y.; Yang, Y.; Duan, Y.; An, Y. LncRNA DCST1-AS1 functions as a competing endogenous RNA to regulate FAIM2 expression by sponging miR-1254 in hepatocellular carcinoma. Clin. Sci. 2019, 133, 367–379. [Google Scholar] [CrossRef]
  137. Tang, L.; Chen, Y.; Tang, X.; Wei, D.; Xu, X.; Yan, F. Long Noncoding RNA DCST1-AS1 Promotes Cell Proliferation and Metastasis in Triple-negative Breast Cancer by Forming a Positive Regulatory Loop with miR-873-5p and MYC. J. Cancer 2020, 11, 311–323. [Google Scholar] [CrossRef]
  138. Liu, J.; Zhang, J.; Hu, Y.; Zou, H.; Zhang, X.; Hu, X. Inhibition of lncRNA DCST1-AS1 suppresses proliferation, migration and invasion of cervical cancer cells by increasing miR-874-3p expression. J. Gene Med. 2021, 23, e3281. [Google Scholar] [CrossRef]
  139. Wang, J.; Lei, C.; Shi, P.; Teng, H.; Lu, L.; Guo, H.; Wang, X. LncRNA DCST1-AS1 Promotes Endometrial Cancer Progression by Modulating the MiR-665/HOXB5 and MiR-873-5p/CADM1 Pathways. Front. Oncol. 2021, 11, 714652. [Google Scholar] [CrossRef]
  140. Su, Y.Z.; Cui, M.F.; Du, J.; Song, B. LncRNA DCST1-AS1 regulated cell proliferation, migration, invasion and apoptosis in gastric cancer by targeting miR-605-3p. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 1158–1167. [Google Scholar] [CrossRef]
  141. Hu, S.; Yao, Y.; Hu, X.; Zhu, Y. LncRNA DCST1-AS1 downregulates miR-29b through methylation in glioblastoma (GBM) to promote cancer cell proliferation. Clin. Transl. Oncol. 2020, 22, 2230–2235. [Google Scholar] [CrossRef] [PubMed]
  142. Huang, L.; Dai, G. Long non-coding RNA DCST1-AS1/hsa-miR-582-5p/HMGB1 axis regulates colorectal cancer progression. Bioengineered 2022, 13, 12–26. [Google Scholar] [CrossRef]
  143. Lv, J.; Li, Q.; Ma, R.; Wang, Z.; Yu, Y.; Liu, H.; Miao, Y.; Jiang, S. Long Noncoding RNA FGD5-AS1 Knockdown Decrease Viability, Migration, and Invasion of Non-Small Cell Lung Cancer (NSCLC) Cells by Regulating the MicroRNA-944/MACC1 Axis. Technol. Cancer Res. Treat. 2021, 20, 1533033821990090. [Google Scholar] [CrossRef] [PubMed]
  144. Zhai, W.; Ye, X.; Wang, Y.; Feng, Y.; Wang, Y.; Lin, Y.; Ding, L.; Yang, L.; Wang, X.; Kuang, Y.; et al. CREPT/RPRD1B promotes tumorigenesis through STAT3-driven gene transcription in a p300-dependent manner. Br. J. Cancer 2021, 124, 1437–1448. [Google Scholar] [CrossRef]
  145. Xia, T.; Zhang, M.; Lei, W.; Yang, R.; Fu, S.; Fan, Z.; Yang, Y.; Zhang, T. Advances in the role of STAT3 in macrophage polarization. Front. Immunol. 2023, 14, 1160719. [Google Scholar] [CrossRef] [PubMed]
  146. Liu, G.; Du, X.; Xiao, L.; Zeng, Q.; Liu, Q. Activation of fgd5-as1 promotes progression of cervical cancer through regulating bst2 to inhibit macrophage m1 polarization. J. Immunol. Res. 2021, 2021, 5857214. [Google Scholar] [CrossRef]
  147. Neil, S.J.; Zang, T.; Bieniasz, P.D. Tetherin inhibits retrovirus release and is antagonized by HIV-1 Vpu. Nature 2008, 451, 425–430. [Google Scholar] [CrossRef]
  148. Kupzig, S.; Korolchuk, V.; Rollason, R.; Sugden, A.; Wilde, A.; Banting, G. Bst-2/HM1.24 is a raft-associated apical membrane protein with an unusual topology. Traffic 2003, 4, 694–709. [Google Scholar] [CrossRef]
  149. Cheng, G.; Liu, D.; Liang, H.; Yang, H.; Chen, K.; Zhang, X. A cluster of long non-coding RNAs exhibit diagnostic and prognostic values in renal cell carcinoma. Aging 2019, 11, 9597–9615. [Google Scholar] [CrossRef]
  150. Deng, Y.; Guo, K.; Tang, Z.; Feng, Y.; Cai, S.; Ye, J.; Xi, Y.; Li, J.; Liu, R.; Cai, C.; et al. Identification and experimental validation of a tumor-infiltrating lymphocytes-related long noncoding RNA signature for prognosis of clear cell renal cell carcinoma. Front. Immunol. 2022, 13, 1046790. [Google Scholar] [CrossRef]
  151. Chureau, C.; Prissette, M.; Bourdet, A.; Barbe, V.; Cattolico, L.; Jones, L.; Eggen, A.; Avner, P.; Duret, L. Comparative sequence analysis of the X-inactivation center region in mouse, human, and bovine. Genome Res. 2002, 12, 894–908. [Google Scholar] [CrossRef]
  152. Sheykhi-Sabzehpoush, M.; Ghasemian, M.; Khojasteh Pour, F.; Mighani, M.; Moghanibashi, M.; Mohammad Jafari, R.; Zabel, M.; Dziegiel, P.; Farzaneh, M.; Kempisty, B. Emerging roles of long non-coding RNA FTX in human disorders. Clin. Transl. Oncol. 2023, 25, 2812–2831. [Google Scholar] [CrossRef]
  153. Liu, Z.; Dou, C.; Yao, B.; Xu, M.; Ding, L.; Wang, Y.; Jia, Y.; Li, Q.; Zhang, H.; Tu, K.; et al. Ftx non coding RNA-derived miR-545 promotes cell proliferation by targeting RIG-I in hepatocellular carcinoma. Oncotarget 2016, 7, 25350–25365. [Google Scholar] [CrossRef]
  154. Rajagopal, T.; Talluri, S.; Akshaya, R.L.; Dunna, N.R. HOTAIR LncRNA: A novel oncogenic propellant in human cancer. Clin. Chim. Acta 2020, 503, 1–18. [Google Scholar] [CrossRef] [PubMed]
  155. Xin, X.; Li, Q.; Fang, J.; Zhao, T. LncRNA HOTAIR: A Potential Prognostic Factor and Therapeutic Target in Human Cancers. Front. Oncol. 2021, 11, 679244. [Google Scholar] [CrossRef] [PubMed]
  156. Wang, J.; Zhao, J.; Hu, P.; Gao, L.; Tian, S.; He, Z. Long Non-coding RNA HOTAIR in Central Nervous System Disorders: New Insights in Pathogenesis, Diagnosis, and Therapeutic Potential. Front. Mol. Neurosci. 2022, 15, 949095. [Google Scholar] [CrossRef] [PubMed]
  157. Zhou, H.; Gao, L.; Yu, Z.H.; Hong, S.J.; Zhang, Z.W.; Qiu, Z.Z. LncRNA HOTAIR promotes renal interstitial fibrosis by regulating Notch1 pathway via the modulation of miR-124. Nephrology (Carlton) 2019, 24, 472–480. [Google Scholar] [CrossRef] [PubMed]
  158. Yu, F.; Chen, B.; Dong, P.; Zheng, J. HOTAIR Epigenetically Modulates PTEN Expression via MicroRNA-29b: A Novel Mechanism in Regulation of Liver Fibrosis. Mol. Ther. 2017, 25, 205–217. [Google Scholar] [CrossRef] [PubMed]
  159. Price, R.L.; Bhan, A.; Mandal, S.S. HOTAIR beyond repression: In protein degradation, inflammation, DNA damage response, and cell signaling. DNA Repair (Amst) 2021, 105, 103141. [Google Scholar] [CrossRef] [PubMed]
  160. Ramya, V.; Shyam, K.P.; Angelmary, A.; Kadalmani, B. Lauric acid epigenetically regulates lncRNA HOTAIR by remodeling chromatin H3K4 tri-methylation and modulates glucose transport in SH-SY5Y human neuroblastoma cells: Lipid switch in macrophage activation. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2024, 1869, 159429. [Google Scholar] [CrossRef]
  161. Zhuang, M.F.; Li, L.J.; Ma, J.B. LncRNA HOTTIP promotes proliferation and cell cycle progression of acute myeloid leukemia cells. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 2908–2915. [Google Scholar] [CrossRef] [PubMed]
  162. Cheng, L.; Yang, Q.; Li, C.; Dai, L.; Yang, Y.; Wang, Q.; Ding, Y.; Zhang, J.; Liu, L.; Zhang, S.; et al. DDA1, a novel oncogene, promotes lung cancer progression through regulation of cell cycle. J. Cell Mol. Med. 2017, 21, 1532–1544. [Google Scholar] [CrossRef] [PubMed]
  163. Cesario, A.; Rocca, B.; Rutella, S. The interplay between indoleamine 2,3-dioxygenase 1 (IDO1) and cyclooxygenase (COX)-2 in chronic inflammation and cancer. Curr. Med. Chem. 2011, 18, 2263–2271. [Google Scholar] [CrossRef] [PubMed]
  164. Sun, P.; Quan, J.C.; Wang, S.; Zhuang, M.; Liu, Z.; Guan, X.; Wang, G.Y.; Wang, H.Y.; Wang, X.S. lncRNA-PACER upregulates COX-2 and PGE2 through the NF-kappaB pathway to promote the proliferation and invasion of colorectal-cancer cells. Gastroenterol. Rep. (Oxf) 2021, 9, 257–268. [Google Scholar] [CrossRef] [PubMed]
  165. Desind, S.Z.; Iacona, J.R.; Yu, C.Y.; Mitrofanova, A.; Lutz, C.S. PACER lncRNA regulates COX-2 expression in lung cancer cells. Oncotarget 2022, 13, 291–306. [Google Scholar] [CrossRef] [PubMed]
  166. Qian, M.; Yang, X.; Li, Z.; Jiang, C.; Song, D.; Yan, W.; Liu, T.; Wu, Z.; Kong, J.; Wei, H.; et al. P50-associated COX-2 extragenic RNA (PACER) overexpression promotes proliferation and metastasis of osteosarcoma cells by activating COX-2 gene. Tumour Biol. 2016, 37, 3879–3886. [Google Scholar] [CrossRef] [PubMed]
  167. Xie, X.; Zhao, W.; Pang, J.; Xiong, X.; Wang, H.; Ma, L. Long non-coding RNA, CHRF, predicts poor prognosis of lung adenocarcinoma and promotes cell proliferation and migration. Oncol. Lett. 2018, 16, 6245–6252. [Google Scholar] [CrossRef]
  168. Wu, W.; Wang, X.; Yu, X.; Lan, H.Y. Smad3 Signatures in Renal Inflammation and Fibrosis. Int. J. Biol. Sci. 2022, 18, 2795–2806. [Google Scholar] [CrossRef]
  169. Luo, S.; Ding, X.; Zhao, S.; Mou, T.; Li, R.; Cao, X. Long non-coding RNA CHRF accelerates LPS-induced acute lung injury through microRNA-146a/Notch1 axis. Ann. Transl. Med. 2021, 9, 1299. [Google Scholar] [CrossRef]
  170. Zeng, H.; Zhou, S.; Cai, W.; Kang, M.; Zhang, P. LncRNA SNHG1: Role in tumorigenesis of multiple human cancers. Cancer Cell Int. 2023, 23, 198. [Google Scholar] [CrossRef]
  171. Li, Z.; Li, X.; Du, X.; Zhang, H.; Wu, Z.; Ren, K.; Han, X. The Interaction Between lncRNA SNHG1 and miR-140 in Regulating Growth and Tumorigenesis via the TLR4/NF-kappaB Pathway in Cholangiocarcinoma. Oncol. Res. 2019, 27, 663–672. [Google Scholar] [CrossRef] [PubMed]
  172. Sheu, K.M.; Hoffmann, A. Functional Hallmarks of Healthy Macrophage Responses: Their Regulatory Basis and Disease Relevance. Annu. Rev. Immunol. 2022, 40, 295–321. [Google Scholar] [CrossRef] [PubMed]
  173. Cai, Y.; Zhang, W.; Xiong, S. Mannose-binding lectin blunts macrophage polarization and ameliorates lupus nephritis. PLoS ONE 2013, 8, e62465. [Google Scholar] [CrossRef] [PubMed]
  174. Schuliga, M. NF-kappaB Signaling in Chronic Inflammatory Airway Disease. Biomolecules 2015, 5, 1266–1283. [Google Scholar] [CrossRef]
  175. Miyata, R.; van Eeden, S.F. The innate and adaptive immune response induced by alveolar macrophages exposed to ambient particulate matter. Toxicol. Appl. Pharmacol. 2011, 257, 209–226. [Google Scholar] [CrossRef]
  176. Zhu, X.; Wang, X.; Wang, Y.; Zhao, Y. The regulatory network among CircHIPK3, LncGAS5, and miR-495 promotes Th2 differentiation in allergic rhinitis. Cell Death Dis. 2020, 11, 216. [Google Scholar] [CrossRef]
  177. Hu, J.; Zhang, L.; Liechty, C.; Zgheib, C.; Hodges, M.M.; Liechty, K.W.; Xu, J. Long Noncoding RNA GAS5 Regulates Macrophage Polarization and Diabetic Wound Healing. J. Invest. Dermatol. 2020, 140, 1629–1638. [Google Scholar] [CrossRef]
  178. Chi, X.; Ding, B.; Zhang, L.; Zhang, J.; Wang, J.; Zhang, W. lncRNA GAS5 promotes M1 macrophage polarization via miR-455-5p/SOCS3 pathway in childhood pneumonia. J. Cell Physiol. 2019, 234, 13242–13251. [Google Scholar] [CrossRef]
  179. Sun, D.; Yu, Z.; Fang, X.; Liu, M.; Pu, Y.; Shao, Q.; Wang, D.; Zhao, X.; Huang, A.; Xiang, Z.; et al. LncRNA GAS5 inhibits microglial M2 polarization and exacerbates demyelination. EMBO Rep. 2017, 18, 1801–1816. [Google Scholar] [CrossRef] [PubMed]
  180. Ito, I.; Asai, A.; Suzuki, S.; Kobayashi, M.; Suzuki, F. M2b macrophage polarization accompanied with reduction of long noncoding RNA GAS5. Biochem. Biophys. Res. Commun. 2017, 493, 170–175. [Google Scholar] [CrossRef] [PubMed]
  181. Xu, Y.; Fang, H.; Xu, Q.; Xu, C.; Yang, L.; Huang, C. LncRNA GAS5 inhibits NLRP3 inflammasome activation-mediated pyroptosis in diabetic cardiomyopathy by targeting miR-34b-3p/AHR. Cell Cycle 2020, 19, 3054–3065. [Google Scholar] [CrossRef] [PubMed]
  182. Lin, G.; Wu, T.; Gao, X.; He, Z.; Nong, W. Research Progress of Long Non-Coding RNA GAS5 in Malignant Tumors. Front. Oncol. 2022, 12, 846497. [Google Scholar] [CrossRef]
  183. Yang, X.; Xie, Z.; Lei, X.; Gan, R. Long non-coding RNA GAS5 in human cancer. Oncol. Lett. 2020, 20, 2587–2594. [Google Scholar] [CrossRef] [PubMed]
  184. Tong, Q.; Gong, A.Y.; Zhang, X.T.; Lin, C.; Ma, S.; Chen, J.; Hu, G.; Chen, X.M. LincRNA-Cox2 modulates TNF-alpha-induced transcription of Il12b gene in intestinal epithelial cells through regulation of Mi-2/NuRD-mediated epigenetic histone modifications. FASEB J. 2016, 30, 1187–1197. [Google Scholar] [CrossRef]
  185. Rapicavoli, N.A.; Qu, K.; Zhang, J.; Mikhail, M.; Laberge, R.M.; Chang, H.Y. A mammalian pseudogene lncRNA at the interface of inflammation and anti-inflammatory therapeutics. eLife 2013, 2, e00762. [Google Scholar] [CrossRef]
  186. Guttman, M.; Amit, I.; Garber, M.; French, C.; Lin, M.F.; Feldser, D.; Huarte, M.; Zuk, O.; Carey, B.W.; Cassady, J.P.; et al. Chromatin signature reveals over a thousand highly conserved large non-coding RNAs in mammals. Nature 2009, 458, 223–227. [Google Scholar] [CrossRef]
  187. Mahmoud, R.H.; Fouad, N.A.; Hefzy, E.M.; Shaker, O.G.; Ahmed, T.I.; Hussein, H.A.; Nasr, M.H.; Zaki, O.M.; Abdelghaffar, N.K.; Abdelaleem, O.O. The potential role of serum expression profile of long non coding RNAs, Cox2 and HOTAIR as novel diagnostic biomarkers in systemic lupus erythematosus. PLoS ONE 2022, 17, e0268176. [Google Scholar] [CrossRef] [PubMed]
  188. Carpenter, S.; Aiello, D.; Atianand, M.K.; Ricci, E.P.; Gandhi, P.; Hall, L.L.; Byron, M.; Monks, B.; Henry-Bezy, M.; Lawrence, J.B.; et al. A long noncoding RNA mediates both activation and repression of immune response genes. Science 2013, 341, 789–792. [Google Scholar] [CrossRef]
  189. Wang, Q.; Xie, Y.; He, Q.; Geng, Y.; Xu, J. LncRNA-Cox2 regulates macrophage polarization and inflammatory response through the CREB-C/EBPbeta signaling pathway in septic mice. Int. Immunopharmacol. 2021, 101, 108347. [Google Scholar] [CrossRef]
  190. Ogulur, I.; Pat, Y.; Ardicli, O.; Barletta, E.; Cevhertas, L.; Fernandez-Santamaria, R.; Huang, M.; Bel Imam, M.; Koch, J.; Ma, S.; et al. Advances and highlights in biomarkers of allergic diseases. Allergy 2021, 76, 3659–3686. [Google Scholar] [CrossRef]
  191. Hussen, B.M.; Azimi, T.; Hidayat, H.J.; Taheri, M.; Ghafouri-Fard, S. NF-KappaB interacting LncRNA: Review of its roles in neoplastic and non-neoplastic conditions. Biomed. Pharmacother. 2021, 139, 111604. [Google Scholar] [CrossRef] [PubMed]
  192. Fernandez-Jimenez, N.; Castellanos-Rubio, A.; Plaza-Izurieta, L.; Irastorza, I.; Elcoroaristizabal, X.; Jauregi-Miguel, A.; Lopez-Euba, T.; Tutau, C.; de Pancorbo, M.M.; Vitoria, J.C.; et al. Coregulation and modulation of NFkappaB-related genes in celiac disease: Uncovered aspects of gut mucosal inflammation. Hum. Mol. Genet. 2014, 23, 1298–1310. [Google Scholar] [CrossRef] [PubMed]
  193. Kaur, P.; Verma, S.; Kushwaha, P.P.; Gupta, S. EZH2 and NF-kappaB: A context-dependent crosstalk and transcriptional regulation in cancer. Cancer Lett. 2023, 560, 216143. [Google Scholar] [CrossRef] [PubMed]
  194. Lai, L.; Wang, G.; Xu, L.; Fu, Y. CEBPB promotes gastrointestinal motility dysfunction after severe acute pancreatitis via the MALAT1/CIRBP/ERK axis. Mol. Immunol. 2023, 156, 1–9. [Google Scholar] [CrossRef] [PubMed]
  195. Cui, Y.; Fu, S.; Sun, D.; Xing, J.; Hou, T.; Wu, X. EPC-derived exosomes promote osteoclastogenesis through LncRNA-MALAT1. J. Cell Mol. Med. 2019, 23, 3843–3854. [Google Scholar] [CrossRef] [PubMed]
  196. Shyu, K.G.; Wang, B.W.; Fang, W.J.; Pan, C.M.; Lin, C.M. Exosomal MALAT1 Derived from High Glucose-Treated Macrophages Up-Regulates Resistin Expression via miR-150-5p Downregulation. Int. J. Mol. Sci. 2022, 23, 1095. [Google Scholar] [CrossRef]
  197. Ye, Y.; Xu, Y.; Lai, Y.; He, W.; Li, Y.; Wang, R.; Luo, X.; Chen, R.; Chen, T. Long non-coding RNA cox-2 prevents immune evasion and metastasis of hepatocellular carcinoma by altering M1/M2 macrophage polarization. J. Cell Biochem. 2018, 119, 2951–2963. [Google Scholar] [CrossRef]
  198. Winkle, M.; El-Daly, S.M.; Fabbri, M.; Calin, G.A. Noncoding RNA therapeutics—challenges and potential solutions. Nat. Rev. Drug Discov. 2021, 20, 629–651. [Google Scholar] [CrossRef]
  199. Arab, I.; Park, J.; Shin, J.J.; Shin, H.S.; Suk, K.; Lee, W.H. Macrophage lncRNAs in cancer development: Long-awaited therapeutic targets. Biochem. Pharmacol. 2023, 218, 115890. [Google Scholar] [CrossRef]
  200. Borgna, V.; Villegas, J.; Burzio, V.A.; Belmar, S.; Araya, M.; Jeldes, E.; Lobos-Gonzalez, L.; Silva, V.; Villota, C.; Oliveira-Cruz, L.; et al. Mitochondrial ASncmtRNA-1 and ASncmtRNA-2 as potent targets to inhibit tumor growth and metastasis in the RenCa murine renal adenocarcinoma model. Oncotarget 2017, 8, 43692–43708. [Google Scholar] [CrossRef] [PubMed]
  201. Lawrence, T.; Fong, C. The resolution of inflammation: Anti-inflammatory roles for NF-kappaB. Int. J. Biochem. Cell Biol. 2010, 42, 519–523. [Google Scholar] [CrossRef] [PubMed]
  202. Oishi, Y.; Spann, N.J.; Link, V.M.; Muse, E.D.; Strid, T.; Edillor, C.; Kolar, M.J.; Matsuzaka, T.; Hayakawa, S.; Tao, J.; et al. SREBP1 Contributes to Resolution of Pro-inflammatory TLR4 Signaling by Reprogramming Fatty Acid Metabolism. Cell Metab. 2017, 25, 412–427. [Google Scholar] [CrossRef] [PubMed]
Figure 1. General action mechanisms of lncRNA. LncRNAs interact with chromatin-remodeling complexes or histone-modifying enzymes to facilitate epigenetic modifications. Their interaction with transcription factors regulates the transcriptional activity of target genes. Interactions with mRNA affect post-transcriptional regulations, including splicing and translation. LncRNAs can also serve as scaffolds for interactions among signaling mediators and various RNAs, including small RNAs, miRNAs and mRNAs. Finally, the miRNA sponging action of lncRNAs may regulate the stability of target mRNAs.
Figure 1. General action mechanisms of lncRNA. LncRNAs interact with chromatin-remodeling complexes or histone-modifying enzymes to facilitate epigenetic modifications. Their interaction with transcription factors regulates the transcriptional activity of target genes. Interactions with mRNA affect post-transcriptional regulations, including splicing and translation. LncRNAs can also serve as scaffolds for interactions among signaling mediators and various RNAs, including small RNAs, miRNAs and mRNAs. Finally, the miRNA sponging action of lncRNAs may regulate the stability of target mRNAs.
Ijms 25 02670 g001
Figure 2. Role of lncRNAs and macrophage NF-κB in the pathogenesis of human diseases. In macrophages, lncRNAs regulate NF-κB activity either directly or indirectly, influencing various processes that lead to inflammation, apoptosis, pyroptosis, and M1/M2 polarization. Additionally, the activation of NF-κB in macrophages affects neighboring cells and the extracellular matrixes, regulating angiogenesis and metastasis in cancer development, as well as plaque rupture in atherosclerosis. These macrophage-associated events are closely linked to chronic inflammation, which ultimately contributes to the progression of various diseases, including sepsis, atherosclerosis, cancer, autoimmunity, and hypersensitivity.
Figure 2. Role of lncRNAs and macrophage NF-κB in the pathogenesis of human diseases. In macrophages, lncRNAs regulate NF-κB activity either directly or indirectly, influencing various processes that lead to inflammation, apoptosis, pyroptosis, and M1/M2 polarization. Additionally, the activation of NF-κB in macrophages affects neighboring cells and the extracellular matrixes, regulating angiogenesis and metastasis in cancer development, as well as plaque rupture in atherosclerosis. These macrophage-associated events are closely linked to chronic inflammation, which ultimately contributes to the progression of various diseases, including sepsis, atherosclerosis, cancer, autoimmunity, and hypersensitivity.
Ijms 25 02670 g002
Figure 3. Mechanisms triggered by macrophage lncRNAs in the regulation of NF-κB activity. Various lncRNAs that regulate the NF-κB activation pathway are depicted. Activating effects are represented by red arrows, while inhibiting effects are shown with blue arrows. LncRNAs that enhance NF-κB activity are indicated in red, while those that suppress NF-κB activity are shown in blue.
Figure 3. Mechanisms triggered by macrophage lncRNAs in the regulation of NF-κB activity. Various lncRNAs that regulate the NF-κB activation pathway are depicted. Activating effects are represented by red arrows, while inhibiting effects are shown with blue arrows. LncRNAs that enhance NF-κB activity are indicated in red, while those that suppress NF-κB activity are shown in blue.
Ijms 25 02670 g003
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

Shin, J.-J.; Park, J.; Shin, H.-S.; Arab, I.; Suk, K.; Lee, W.-H. Roles of lncRNAs in NF-κB-Mediated Macrophage Inflammation and Their Implications in the Pathogenesis of Human Diseases. Int. J. Mol. Sci. 2024, 25, 2670. https://doi.org/10.3390/ijms25052670

AMA Style

Shin J-J, Park J, Shin H-S, Arab I, Suk K, Lee W-H. Roles of lncRNAs in NF-κB-Mediated Macrophage Inflammation and Their Implications in the Pathogenesis of Human Diseases. International Journal of Molecular Sciences. 2024; 25(5):2670. https://doi.org/10.3390/ijms25052670

Chicago/Turabian Style

Shin, Jae-Joon, Jeongkwang Park, Hyeung-Seob Shin, Imene Arab, Kyoungho Suk, and Won-Ha Lee. 2024. "Roles of lncRNAs in NF-κB-Mediated Macrophage Inflammation and Their Implications in the Pathogenesis of Human Diseases" International Journal of Molecular Sciences 25, no. 5: 2670. https://doi.org/10.3390/ijms25052670

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