Next Article in Journal
Sustained Microglial Activation Promotes Synaptic Loss and Neuronal Dysfunction after Recovery from ZIKV Infection
Previous Article in Journal
Increased Expression of the Neuropeptides PACAP/VIP in the Brain of Mice with CNS Targeted Production of IL-6 Is Mediated in Part by Trans-Signalling
Previous Article in Special Issue
The Exosomes of Stem Cells from Human Exfoliated Deciduous Teeth Suppress Inflammation in Osteoarthritis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Functions and Clinical Applications of Extracellular Vesicles in TH2 Cell-Mediated Airway Inflammatory Diseases: A Review

1
Department of Biomedical Science, Korea University College of Medicine, Seoul 02841, Republic of Korea
2
Department of Otorhinolaryngology-Head & Neck Surgery, Korea University College of Medicine, Seoul 02841, Republic of Korea
3
Neuroscience Research Institute, Korea University College of Medicine, Seoul 02841, Republic of Korea
4
Mucosal Immunology Institute, Korea University College of Medicine, Seoul 02841, Republic of Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(17), 9455; https://doi.org/10.3390/ijms25179455
Submission received: 29 July 2024 / Revised: 24 August 2024 / Accepted: 28 August 2024 / Published: 30 August 2024

Abstract

:
Type 2 airway inflammation (T2AI), driven by type 2 innate lymphoid and CD4+ T helper 2 cells, leads to various diseases and conditions, such as chronic rhinosinusitis with nasal polyps, allergic rhinitis, and asthma. Emerging evidence suggests the involvement of extracellular vesicles (EVs) in these diseases. In this review, we describe the immunological T2AI pathogenic mechanisms, outline EV characteristics, and highlight their applications in the diagnosis and treatment of T2AI. An extensive literature search was conducted using appropriate strategies to identify relevant articles from various online databases. EVs in various biological samples showed disease-specific characteristics for chronic rhinosinusitis with nasal polyps, allergic rhinitis, and asthma, with some demonstrating therapeutic effects against these conditions. However, most studies have been limited to in vitro and animal models, highlighting the need for further clinical research on the diagnostic and therapeutic applications of EVs.

1. Introduction

Type 2 airway inflammation (T2AI) is characterized by an immune response mediated by type 2 innate lymphoid cells (ILC2s) and a subset of CD4+ T helper 2 (TH2) cells in the respiratory system. The “one-airway-one-disease” hypothesis postulates that the upper and lower airways share a similar pathophysiology [1]. Therefore, excessive type 2 inflammation (T2I) is able to lead to chronic rhinosinusitis (CRS), allergic rhinitis (AR), and asthma depending on its location in the airway tract. Over the past 5 years, the global prevalence of type 2 airway diseases has steadily increased, with CRS, AR, and asthma affecting approximately 5–12% [2], 10–30% [3], and 3–10% [4] of the global population, respectively. These diseases exhibit correlated characteristics under airway inflammatory conditions [5]. For example, individuals with asthma often exhibit AR and CRS symptoms [6], whereas those with CRS commonly exhibit asthma as a comorbidity [7]. Despite advancements in diagnosis and medication, treatment remains challenging, attributable to the complexity of these comorbid conditions, which share a similar pathogenesis in the upper and lower airway tracts [1,2,8,9]. The overlapping symptoms and inflammatory pathways in CRS, AR, and asthma further complicate their diagnosis and treatment, necessitating additional research and improved clinical strategies for managing these conditions.
Recent studies have increasingly highlighted the diagnostic and therapeutic applications of extracellular vesicles (EVs) for various diseases [10,11,12]. EVs, vesicles derived from cellular membranes, transport biomacromolecules to recipient cells. As EVs act as carriers of cargo molecules, they are considered important mediators of cellular communication [13]. Owing to their heterogeneity, which is influenced by the conditions of the secreting cells and their role in delivering cargo molecules to specific targets, EVs are considered promising agents for the diagnosis and therapy of various diseases [14].
In this review, we discuss the intricate pathogenesis of T2AI and outline the characteristics of EVs. Furthermore, we highlight the current potential applications of EVs for the diagnosis and treatment of T2AI-related diseases. We summarize recent research findings that demonstrate the potential of EVs as biomarkers for early disease detection and as therapeutic agents capable of modulating immune responses and inhibiting inflammation. We provide a comprehensive overview of the functions and clinical applications of EVs in T2AI diseases and outline future research directions to facilitate and advance effective diagnostic and therapeutic strategies.

2. The Pathogenesis of T2AI Diseases

The airway is anatomically divided into the upper respiratory tract, including the nasal cavity, pharynx, and larynx, and the lower respiratory tract, including the trachea, bronchi, and bronchioles [15]. Excessive T2I predominantly causes CRS and AR in the paranasal sinuses and turbinates, whereas asthma primarily affects the lower respiratory tract. The airway is a tract largely composed of mucosa, a tissue comprising various epithelial cells that primarily regulate immune responses by serving as physical barriers [16]. Various genetic and environmental factors can stimulate this epithelial layer, leading to barrier disruption and induction of T2I [17,18].

2.1. T2I Occurs via Interactions among the Epithelial Cells, Dendritic Cells, and T-Cells

Disrupted epithelial cells release chemokines, including C-C motif chemokine ligands (CCLs), and alarmin cytokines, such as the thymic stromal lymphopoietin (TSLP), interleukin (IL)-25, and IL-33 [18,19]. These chemokines and cytokines not only activate ILC2s but also attract dendritic cells (DCs) expressing C-C motif chemokine receptors to the site of disruption. At this location, immature DCs (imDCs) are converted into mature DCs (mDCs) via stimulation by alarmin cytokines and the binding of antigens and allergens to pattern recognition receptors on the surface membrane of DCs [18]. This maturation process is crucial for the initiation and propagation of immune responses, since mDCs function as key antigen-presenting cells [20]. They express major histocompatibility complex class II (MHC II) molecules, and the processed antigens or allergen peptide fragments are presented on these molecules. This presentation is essential for T-cell activation and adaptive immune responses [21]. Activated mDCs migrate to the local lymph nodes, where they polarize naïve T cells into TH2 cells by forming an immune synapse with T-cells. The binding of antigen or allergen peptide fragment-loaded MHC II to the T-cell receptor (TCR) provides the primary signal necessary for T-cell polarization [22]. This interaction leads to the differentiation and expression of antigen- or allergen-specific TH2 cells, which are essential for adaptive immune responses [23].

2.2. TH2 Cells Trigger Adaptive Immune Responses by Interacting with B-Cells

Polarized TH2 cells in the lymph nodes migrate to the mucosa, where they are activated by alarmin cytokines derived from epithelial cells and ILC2s. Once activated, both ILC2s and TH2 cells release the signature TH2 cytokines, IL-4, IL-5, and IL-13 [24]. These cytokines play critical roles in immune responses by inducing B-cells to undergo immunoglobulin E (IgE) class switching during their interactions with TH2 cells [25]. Activated B-cells differentiate into plasma cells, which secrete IgE and memory B-cells, facilitating a rapid adaptive immune response to specific antigens or allergens [26].

2.3. T2I Is Initiated by Innate Immunity Activation by TH2 and B-Cells

TH2 cytokines secreted by ILC2s and TH2 cells affect epithelial cells, granulocytes, and B-cells. IL-4 induces the transformation of epithelial cells into goblet cells, which primarily produce the mucus [27]. IL-5 activates granulocytes, particularly eosinophils, leading to their degranulation [28]. The stimulated eosinophils degranulate, causing mucus thickening and increasing epithelial damage [27,29].
IgE released from plasma cells binds to the IgE receptor (FcεR) on mast cell membranes through a process known as sensitization in allergic responses. As IgE bound to mast cells is specific to certain antigens or allergens, when these antigens or allergens re-enter the disrupted epithelial cells, they directly bind to the IgE on mast cells. This cross-linking triggers mast cell degranulation, releasing inflammatory molecules, such as histamines, prostaglandins, and leukotrienes [30]. These factors ultimately induce inflammatory symptoms, such as coughing, sneezing, and airway congestion [17]. Figure 1 illustrates the overall immunological pathogenic mechanism of T2AI.

3. EVs

3.1. The Biogenesis of EVs

EVs are secreted by donor cells through various biological processes and they are primarily classified into apoptotic bodies (ApoBDs), micro-vesicles, and exosomes based on their cell of origin and vesicle size [31,32,33]. ApoBDs (diameter: 1–5 µm) are vesicles shed from apoptotic cells during programmed cell death. They contain proteins, lipids, and genetic materials, such as mRNAs, micro-RNAs (miRNAs), and other cellular organelles [31]. These vesicles are formed by a tightly regulated process, called apoptotic cell disassembly, which involves plasma membrane blebbing, formation of apoptotic protrusions, and fragmentation into ApoBDs [34]. ApoBDs are produced by dying cells, whereas micro-vesicles and exosomes are formed via the budding or exocytosis of living cells.
Micro-vesicles are formed by the outward budding of the cellular membrane, leading to the detachment and formation of vesicles larger than 200 nm in diameter. Membrane budding occurs at specific sites on the plasma membrane and is influenced by phospholipid redistribution. This mechanical redistribution, along with the phosphorylation of ρ-kinase-mediated myosin light chain and contractile machinery, facilitates the pinching and detachment of vesicles [35,36]. Unlike ApoBDs, micro-vesicles do not contain cellular organelles but do include proteins, lipids, and genetic materials. The cytosolic biomolecules are incorporated randomly, whereas membrane proteins and receptors are specifically directed to the plasma membrane before micro-vesicle budding. Micro-vesicles might be identified by their lipid composition, plasma membrane receptors, and biomolecular content, which reflect their cellular origin [32].
Similar to micro-vesicles, exosomes are vesicles that contain proteins, lipids, and genetic materials; however, they have a more complex and heterogeneous formation mechanism than micro-vesicles and are typically less than 150 nm in diameter. They are mainly classified into classical and non-classical exosomes (exosome-like vesicles). Classical exosomes are derived from the endosomal sorting complex required for the transport (ESCRT)-dependent pathway. This pathway involves the inward invagination of the endosomal membrane facilitated by ESCRT proteins, such as ESCRT-0 (hepatocyte receptor tyrosine kinase substrate [HRS], also known as the vacuolar protein sorting-associated protein [VPS]-27), ESCRT-I (tumor susceptibility gene 101 [TSG101]), ESCRT-II (VPS25), and ESCRT-III (charged multivesicular body protein [CHMP]-4A, CHMP4B, and CHMP4C) [37]. These proteins recruit accessory proteins, such as the ALG-2-interacting protein X, leading to the formation of multivesicular bodies (MVBs) that contain several intraluminal vesicles (ILVs) [38]. MVBs are capable of being formed via an ESCRT-independent pathway involving sphingomyelinase hydrolysis and ceramide formation [39]. During MVB formation, ILVs package endosomal or cytosolic proteins and genetic materials that are reflective of their parental cells. Subsequently, tetraspanins proteins, such as CD9, CD63, and CD81, are incorporated into MVBs, as they are associated with recycling routes between the plasma membrane and various cellular organelles [40]. MVBs are produced via either ESCRT-dependent or-independent mechanisms and subsequently fuse with the cellular membrane to release ILVs, which become exosomes or exosome-like vesicles [41].
Synthesized and released from donor-cells, EVs are capable of inducing various intra-cellular alterations of recipient cells. For example, When RNA contained within EVs is delivered to target cells, it contributes to gene expression, leading to changes in protein synthesis within the recipient cells [42,43]. Also, EVs carrying miRNAs are able to suppress the expression of their target genes within recipient cells [44].

3.2. The Classification of EVs

EVs are primarily classified into ApoBDs, micro-vesicles, and exosomes. Accurately categorizing EVs is challenging because of their potential to carry markers common to multiple biogenic pathways prior to secretion. The International Society for Extracellular Vesicles has provided comprehensive guidelines on characterizing EVs and assessing their purity [45]. The Minimal Information for Studies of EVs 2018 recommends the use of recent EV nomenclature to categorize EVs based on various criteria, including physical properties, specific markers, and the conditions or cells of origin. This approach aims to standardize EV research and improve the reproducibility and accuracy of findings in the field. Figure 2 illustrates the overall process of the generation and classification of EVs.

3.3. The Detection and Analysis of EVs

Recently, several techniques for detecting and analyzing EVs from biological samples have been developed, including ultracentrifugation, size-exclusion chromatography, nanoparticle tracking analysis (NTA), flow cytometry, and Western blotting (WB). Since each separation method has its own unique advantages and limitations, it is crucial to understand their specific characteristics and select the most suitable technique for the sample at hand to ensure highly reproducible results.

4. The Applications of EVs for Various T2AI Diseases

Among the various EVs subtypes, exosomes or small EVs (sEVs) are the most commonly utilized and applied in various diseases [46]. Similar to exosomes, sEVs measure between 30 and 150 nm in diameter and constitute the most prevalent subset of EVs in biological fluids [47]. In the case of T2AI diseases, the majority of research has also focused on the use of exosomes or sEVs.

4.1. The Application of EVs for CRS

CRS is a complex inflammatory disease affecting the mucosa of the nasal cavity and paranasal sinuses in the upper airway tract. Common symptoms of CRS include facial pain or pressure, nasal discharge, congestion, and a reduced sense of smell (hyposmia) or a complete loss of smell (anosmia). It significantly impacts quality of life and poses a substantial economic burden on affected patients. Traditionally, CRS is categorized based on the presence/absence of nasal polyps as CRSsNPs (without nasal polyps) and CRSwNPs (with nasal polyps) [2,48]. Patients with CRSwNP are more likely to experience concurrent lower airway inflammation, such as asthma, than those with CRSsNP. Compared with those with CRSsNP, patients with CRSwNP typically exhibit more severe symptoms, require more frequent surgeries, and have higher recurrence rates after surgical intervention. These factors are crucial considerations for their effective clinical treatment [49]. CRSwNP arises from heterogeneous conditions and is primarily classified into type 2 and non-type 2 endotypes. Type 2 CRSwNP, common in Western countries, is mediated by T2I, whereas non-type 2, including types 1 and 17, is mediated by TH1 and TH17 cells and their signature cytokines, interferon (IFN)-γ and IL-17, respectively [50]. Recently, CRSwNP has been further classified into eosinophilic CRSwNP (eCRSwNP) and non-eCRSwNP based on the histological quantification of eosinophil numbers [51]. ECRSwNP is strongly associated with T2I and worse clinical prognosis than non-eCRSwNP [52,53]. Consequently, many studies have focused on CRSwNP mediated by T2I or eCRSwNP to improve patient treatment outcomes.
Recently, numerous studies have focused on detecting altered biomacromolecules in patients with CRSwNP compared to biomolecules in the normal groups using EVs derived from various biological samples. In the mucus, expression of cystatin (CST)-2, also known as CST-SA, is significantly higher in the nasal mucus-derived EVs (NMDEs) of patients with CRSwNP than in those of patients with CRSsNP and controls. Moreover, CST-2 expression is closely related to tissue eosinophilia, asthma, and allergies [54]. This finding is supported by previous studies indicating that CSTs, which are endogenous protease inhibitors, play crucial roles in maintaining the epithelial barrier and modulating immune responses. CSTs also contribute to the recurrence of eosinophilic CRS by interacting with various components, including TH2 cytokines and fibroblasts [55,56]. Mueller et al. reported that the expression of specific serpin family members, including SerpinB2, SerpinE1, SerpinF2, and SerpinG1, are up-regulated in both the matched tissues (CRSwNP: middle turbinate polypoid mucosa; control: inferior turbinate) and NMDEs of patients with CRSwNP compared with their expression in the controls [57]. These serpins inhibit the fibrinolysis pathway by inactivating the tissue plasminogen activator, urokinase plasminogen activator, and plasmin. Therefore, increased expression of the serpin family in CRSwNP possibly contributes to polyp formation by maintaining fibrin deposition [57]. Moreover, pappalysin-A (PAPP-A) levels were significantly increased in both the matched tissues (CRSwNP: adjacent nasal polyp/control: middle turbinate) and NMDEs of patients with CRSwNP. PAPP-A cleaves its substrates, such as insulin-like growth factor (IGF) binding proteins (IGFBP)-4/5, which inactivate IGFs, thereby leading to the activation of IGFs. Therefore, up-regulation of the PAPP-A/IGFBP-4/5/IGF-1 axis in CRSwNP is possibly associated with enhanced epithelial proliferation and polyp growth [58].
Differential expression of specific targets in CRSwNP has been reported in blood samples. He et al. reported that patients with CRSwNP exhibit differentially expressed miRNAs (DEMs) in plasma, with increased levels of miR-79, miR-677, and miR-1037 and decreased levels of miR-4, miR-192, and miR-1022 compared to the controls [59]. miRNAs are non-coding RNAs, approximately 21–24 nucleotides long, that play post-transcriptional regulatory roles by binding to target mRNAs, leading to mRNA degradation or the inhibition of protein translation [60]. These DEMs in CRSwNP plasma are involved in various pathways, such as extracellular matrix–receptor interactions, calcium signaling, and the Hippo, Notch, ErbB, and cAMP signaling pathways [54]. Moreover, galectin 10 (GAL10) and eosinophil peroxidase (EPO) levels in both the tissues (CRSwNP: nasal polyp/CRSsNP: inferior turbinate) and serum EVs are higher in patients with CRSwNP than in those with CRSsNP [61]. Galectins are a family of 15 lectins implicated in various biological processes, including immune responses, inflammation, cell proliferation, motility, and programmed cell death [62]. EPO, a heme protein that is abundantly expressed in eosinophils, catalyzes the production of cytotoxic oxidants and contributes to the pathogeneses of cancer, asthma, and allergic inflammatory disorders [63]. Notably, GAL10 in the EVs released from eosinophils is possibly associated with the development of CRSwNP by promoting eosinophil extracellular trap cell death, which leads to sinus obstruction [61].
EVs of patients with CRSwNP are able to be detected and analyzed using biological fluids, such as mucus and blood samples, and primary cells. Zhou et al. demonstrated that the primary human nasal epithelial cells (hNECs) of patients with CRSwNP contain differentially expressed proteins (DEPs) in their derived exosomes. These patients exhibit increased levels of integrin-β5, endoplasmic reticulum membrane protein complex subunit 4, L1 cell adhesion molecule, and transaldolase 1 and decreased levels of torsin family 4 member A, small proline-rich protein 3, and solute carrier family 1 member 1 compared to the healthy controls. These DEPs are mainly involved in epithelial remodeling. Notably, exosomes derived from the epithelial cells of patients with CRSwNP significantly inhibit the proliferation rate of normal hNECs [64]. The inhibition of adjacent normal cells by DEPs in EVs derived from the hNECs of patients with CRSwNP may be due to the activation of the p53 signaling pathway, leading to the suppression of proliferation [64]. Table 1 provides a detailed overview of EVs investigated in recent CRSwNP studies.

4.2. The Application of EVs for AR

AR occurs via IgE-mediated responses to inhaled allergens, leading to mucosal inflammation driven by T2I in the upper airway. It is characterized by various symptoms, such as sneezing, nasal congestion, nasal itching, and rhinorrhea (nasal discharge). AR is a prevalent chronic condition, especially in high-income countries, where its prevalence can reach up to 50%, but it is relatively rare in low- and middle-income countries [65]. AR is a significant global health concern that contributes substantially to the overall burden of disease and disability worldwide. Moreover, AR is associated with decreased productivity at work and school, sleep disturbance, and reduced participation in outdoor activities among children [66,67].
Table 1. List of studies investigating extracellular vesicles (EVs) in chronic rhinosinusitis with nasal polyps (CRSwNP).
Table 1. List of studies investigating extracellular vesicles (EVs) in chronic rhinosinusitis with nasal polyps (CRSwNP).
No.EV DonorUsed EVsEV Isolation
Methods
EV Identification MethodsSubjectsResults/EffectsRef.
1MucusExosomes/sEVsUltracentrifugationNot availablePatients with
CRSwNP vs. Control
CST-2 (CST-SA) ↑[54]
2MucusExosomes/sEVsCentrifugation (1500× g for 30 min, 12,000× g for 45 min)
Ultracentrifugation (110,000× g for 2 h) Filtration with 0.22-µm filter
Centrifugation
(110,000× g for 70 min)
Not availablePatients with
CRSwNP vs. Control
SerpinB2, SerpinE1, SerpinF2, SerpinG1 ↑[57]
3MucusExosomes/sEVsCentrifugation
(1500× g for 30 min, 12,000× g for 45 min)
Ultracentrifugation
(110,000× g for 2 h)
Filtration with 0.22-µm filter
Centrifugation
(110,000× g for 70 min)
Not availablePatients with
CRSwNP vs. Control
PAPP-A ↑[58]
4PlasmaExosomes/sEVsCentrifugation
(5000× g for 20 min)
Filtration with 0.45-µm filter
Size exclusion column
NTA (100 nm)
WB (CD9, CD63, TSG101, ALIX)
TEM (40–160 nm)
Patients with
CRSwNP vs. Control
miR-677, miR-1037, miR-79 ↑
miR-192, miR-1022, miR-4 ↓
[59]
5SerumExosomes/sEVsPS affinity methodTEM (CD9)
NTA (<200 nm)
Patients with
CRSwNP vs. CRSsNP
GAL10, EPO ↑[61]
6Epithelial cellExosomes/sEVsCentrifugation
(1000× g for 10 min, 16,500× g for 30 min)
Ultracentrifugation
(100,000× g for 2 h)
NTA
TEM
WB (CD9, TSG101)
Patients with
CRSwNP vs. Control
EV treatment after vs. before
into normal hNEC
ITGB5, EMC4, L1CAM, TALDO1 ↑
TOR4A, SPRR3, SLC1A1 ↓
Rate of proliferation ↓
[64]
A vs. B: Relative results of group A group vs. group B. Abbreviations: No., Number; EVs, extracellular vesicles; CRSwNP, chronic rhinosinusitis with nasal polyp; CRSsNP, chronic rhinosinusitis without nasal polyp; sEVs, small EVs; PS, phosphatidylserine; NTA, nanoparticle tracking analysis; WB, western blotting; TEM, transmission electron microscopy; CD, cluster of differentiation; TSG, tumor susceptibility gene; ALIX, ALG-2 interacting protein X; hNEC, human nasal epithelial cell; CST-2, cystatin-2; PAPP-A, pappalysin-A; miR, micro-RNA; GAL10, galectin 10; EPO, eosinophil peroxidase; ITGB5, integrin-β5; EMC4, endoplasmic reticulum membrane protein complex subunit 4; TALDO1, transaldolase 1; TOR4A, torsin family 4 member A; SPRR3, small proline-rich protein 3; SLC1A1, solute carrier family 1 member 1; Ref., reference. ↑: Up-regulation, ↓: Down-regulation.
As shown in Table 2, recently, many studies have been focusing on identifying biomacromolecules altered in individuals with AR, specifically in EVs derived from various biological specimens, compared to those in the normal cohorts. In the nasal mucus, the NMDEs of patients with AR exhibit DEMs, with up-regulated miR-30-5p, miR-199b-3p, and miR-203 levels and down-regulated miR-28-3p, miR-874, and miR-875-5p levels. These DEMs are associated with B-cell receptors and salivary secretion signaling pathways [68]. Li et al. reported low levels of miR-146a-5p in nasal epithelial tissues, NMDEs, and hNECs in patients with AR. As miR-146a-5p promotes TH1 differentiation by inhibiting the interaction between mothers against decapentaplegic homolog 3 (SMAD3) and GATA-binding protein 3 (GATA3), its reduced expression is closely related to AR [69]. Similarly, Zhu et al. presented differentially expressed non-coding RNA in patients with AR using both NMDEs and hNEC-derived EVs. Long non-coding RNA GAS5 (lncGAS5) levels are up-regulated in the NMDEs of patients with AR and OVA-treated hNEC-derived exosomes compared to those in the controls [70]. lncRNAs are a subgroup of non-coding RNAs characterized by lengths exceeding 200 nucleotides [71]. Previous studies have partially unveiled the regulatory roles of lncRNAs in influencing the immune system [72]. Increased expression of lncGAS5 induces TH2 cell differentiation and AR by decreasing the expression of T-bet and IFN-γ and increasing the expression of GATA3 and IL-4 in CD4+ T-cells co-incubated with NMDEs from patients with AR compared to those in the cells co-incubated with NMDEs from controls [70]. As the TH1 signature cytokines, including IL-2 and IFN-γ, and TH2 signature cytokines are antagonistic toward each other [73], regulation of the TH1/TH2 ratio is important for AR management [74]. Differential expression of specific molecules in AR has been reported in blood samples. EVs obtained from the plasma of patients with AR exhibit increased expression of the peptide fragments of Der p1—a protease from the house mite Dermatophagoides pteronyssinus—which is notably related to the symptoms of patients with AR and plasma IL-13 levels by inducing the differentiation of TH2 cells [75]. The Der p1, which is commonly found in house dust mites, disrupts the epithelial barrier [76]. Therefore, EVs containing Der p1 may be closely related with T2I pathogenesis, especially during antigen presentation.
Unlike the limited research on CRSwNPs, many studies have assessed the potential therapeutic applications of EVs for AR. Mesenchymal stromal cell (MSC)-derived small EVs (sEVs) are promising therapeutic agents for AR. In a previous study, treatment of MSC-derived sEVs delivering miR-146a-5p into ILC2s derived from the peripheral blood mononuclear cells of patients with AR decreased IL-4, IL-5, and IL-13 levels and increased IL-1α, IL-1β, IL-6, and IL-1 receptor antagonist (IL-1RN) levels [77]. This aligns with previous reports that miR-146a-5p levels are decreased in the NMDEs of patients with AR and supports the immunoregulatory effects of MSCs against various immunological diseases [69]. MiR-146a-5p has been identified as a molecule that regulates IL-33 release from ILC2s by inhibiting IRAK1 (IL-1R-associated kinase 1) and TRAF6 (Tumor necrosis factor receptor-associated factor 6), which are key regulators in the IL-33-ILC2 signaling pathway [78]. Therefore, MSC-derived EVs containing miR-146a-5p may have a beneficial effect on AR.
MSC-derived sEVs do not just affect DCs and T-cells; ILC2s ImDCs treated with MSC-derived sEVs exhibit reduced levels of CD11c, CD40, CD80, CD86, and HLA-DR, inhibiting their maturation compared to pre-sEV treated DCs. Furthermore, mDCs treated with these sEVs decrease the counts of IL-4+ and IL-13+ TH cells and increase the counts of regulatory T-cells (Tregs) by promoting IL-10+ TH cells through co-culture with mDCs and CD4+ T-cells from the peripheral blood mononuclear cells of patients with AR [79]. Tregs inhibit TH2 cell functions by producing IL-10 [80]. These reports suggest that MSC-derived sEVs could be promising therapeutic agents for AR.
Table 2. List of studies investigating extracellular vesicles (EVs) in allergic rhinitis (AR).
Table 2. List of studies investigating extracellular vesicles (EVs) in allergic rhinitis (AR).
No.EV DonorUsed EVsEV Isolation
Methods
EV Identification MethodsSubjectsResults/EffectsRef.
1MucusExosomes/sEVsCentrifugation
(3000× g for 15 min, 10,000× g for 30 min, 50,000× g for 1 h, 100,000× g for 1 h)
Bead-based flow cytometry
(MHCII, CD63)
AR patients vs. ControlmiR-30-5p, 199b-3p, and 203 ↑
miR-28-3p, 874, and 875-5p ↓
[68]
2Nasal mucusExosomes/sEVs Centrifugation
(12,000× g for 45 min at 4 °C)
Supernatant centrifugation
(110,000× g for 2 h)
Filtration with 0.22-μm filter
Centrifugation
(110,000× g for 70 min)
TEM, NTA (10–210 nm),
WB (CD9, CD63, ALIX)
AR patients vs. ControlmiR-146a-5p ↓[69]
3Mucus/Epithelial cellExosomes/sEVsCentrifugation
(12,000× g for 45 min)
Ultracentrifugation
(110,000× g for 2 h)
Filtration with a 0.22-μm filter
Centrifugation
(110,000× g for 70 min)
TEM,
WB (CD63, CD81)
AR patients vs. Control OVA treatment vs. non-treatment in RPMI-2650
AR patient EVs vs. Control EVs co-incubated with CD4+ T-cell
LncGAS5 ↑
IFN-γ, T-bet ↓ in CD4+ T-cells
IL-4, GATA3 ↑ in CD4+ T-cells
[70]
4PlasmaUnknownUltracentrifugationNTA, TEM,
WB (CD9, CD63, CD81, ALIX, TSG101)
AR patients vs. ControlDer p1 ↑[75]
5MSCssEVsCentrifugation
(2000× g for 20 min)
Anion-exchange chromatography
Ultracentrifugation
(300× g for 5 min, 2000× g for 20 min, 12,000× g for 30 min, 110,000× g for 70 min)
Flow cytometry
(CD9, CD63, CD81),
ELISA (CD63),
WB (CD9, CD63, CD81, ALIX, TSG101) Protein concentration, NTA, TEM
EV treatment after vs. before
using ILC2s from patients with AR PBMCs
IL-4, IL-5, and IL-13 ↓
IL-1α, IL-1β, IL-6, and IL-1RN ↑ ILC2 function ↓
[77]
6MSCssEVsCentrifugation
(2000× g for 20 min)
Protein concentration
NTA (85–284 nm)
TEM (<150 nm)
WB (CD9, CD63, ALIX, TSG101)
EV treatment after vs. before using imDCs
EV-treated DCs vs. Control DCs co-cultured with T-cells from patients with AR PBMCs
CD11c, HLA-DR, CD40, CD80, and CD86 ↓
IL-4 and IL-13 ↓ in T-cells
IL-10 ↑ in T-cells
Treg expansion ↑
[79]
7Hypoxic-MSCsHypoxic-MSC-derived EVsUltracentrifugation
(100,000× g for 90 min)
TEM, NTA,
WB
(CD63, CD9, TSG101)
Treatment with hypoxic EVs vs. normal EVs
of AR mice
EV treatment after vs. before using imDCs
VEGF ↑ in hypoxic EVs
IL-4, IL-10, mucosa thickness, and inflammation ↓
in nasal mucosa
CD40, CD80, and CD83 ↓
[81]
8hADSCsUnknownMedium Centrifugation
(300× g for 15 min, 4000× g for 15 min, 10,000× g for 30 min)
Ultracentrifugation
(130,000× g for 90 min)
TEM, NTA (100 nm),
WB
(CD63, CD81, HSP70)
EV treatment after vs. before
in AR mice
Nasal symptoms and inflammatory cell infiltration ↓
IgE, IL-4, and IFN-γ ↓ in serum
Ratio of TH1/TH2 ↑
[82]
A vs. B: Relative results of group A vs. group B. Abbreviations: No., Number; EVs, extracellular vesicles; AR, allergic rhinitis; MSCs, mesenchymal stromal cells; hADSCs, human adipose tissue-derived stem cells; sEVs, small EVs; TEM, transmission electron microscopy; NTA, nanoparticle tracking analysis; WB, western blotting; MHC, major histocompatibility complex; CD, cluster of differentiation; ALIX, ALG-2 interacting protein X; TSG, tumor susceptibility gene; HSP, heat-shock protein; ILC2, type 2 innate lymphoid cell; DC, dendritic cell; imDC, immature DC; mDC, mature DC; PBMC, peripheral blood mononuclear cell; miR, micro-RNA; Lnc, long non-coding RNA; IFN, interferon; IL, interleukin; GATA, GATA binding protein; Der p1, Dermatophagoides pteronyssinus 1; IL-1RN, IL-1 receptor antagonist; VEGF, vascular endothelial growth factor; IgE, immunoglobulin E; TH, helper T-cell; Ref., reference. ↑: Up-regulation, ↓: Down-regulation.
The effects of MSC-derived EVs on AR have been demonstrated in vivo using an AR mouse model. Notably, EVs derived from MSCs cultured under hypoxic conditions exert more pronounced therapeutic effects than those cultured under non-hypoxic and pre-treatment conditions. This enhanced effect is attributed to the inhibition of imDC maturation and reduction in IL-4 and IL-5 expression, nasal mucosa thickness, and T2I in the nasal mucosa, which are largely mediated by the vascular endothelial growth factor present in EVs [81]. In addition to MSCs, human adipose tissue-derived stem cells (hADSCs) exhibit promising effects against AR. In a previous study, similar to MSCs, hADSCs-derived EVs decreased the serum levels of soluble IgE, IFN-γ, and IL-4, ultimately reducing inflammatory cell infiltration and AR symptoms by enhancing the TH1/TH2 ratio in an AR mouse model [82].

4.3. The Application of EVs for Asthma

Asthma is a complex respiratory condition characterized by inflammation of the lower airway tract, bronchial hyperresponsiveness, and variable airflow obstruction [83]. It is often associated with other complications, such as AR and CRSwNP [6]. Typical symptoms of asthma include wheezing, shortness of breath (dyspnea), chest tightness, and coughing. Asthma is an inflammatory disease exhibiting diverse characteristics and heterogeneous phenotypes and endotypes similar to those observed in CRSwNP. Notably, eosinophilic asthma (eA), predominantly mediated by T2AI, is the most common phenotype/endotype of severe asthma [84].
Many studies have investigated the roles of EVs in asthma (Table 3). Preliminary studies have investigated the effects of IL-13 on normal human bronchial epithelial cells (hBECs) under asthmatic conditions in vitro. One study reported increased exosome secretion in healthy hBECs after IL-13 treatment compared to the pre-treatment levels, facilitating monocyte proliferation and chemotaxis [85]. Another study using healthy primary hBECs reported decreased expression levels of miR-34a, miR-92b, and miR-210 after IL-13 treatment. The target genes of these miRNAs are mainly involved in DC maturation and regulation of TH2 differentiation [86]. Additionally, DEMs, such as let-7, miR-9, and miR-10, have been identified in the hBECs of patients with asthma. These miRNAs are primarily associated with TCR signaling pathways, T-cell differentiation, and plasma cell differentiation [87]. These reports suggest that EVs and their miRNA cargos play significant roles in the pathogenesis and progression of asthma by modulating key immune processes.
Another study investigated EVs under in vitro asthmatic conditions using normal hBECs, particulate matter less than 2.5 µm in diameter (PM2.5), and blood samples from patients with asthma. The study reported that let-7i-5p levels are significantly up-regulated in both normal hBECs treated with PM2.5 and plasma-derived EVs from patients with asthma compared to those in the pre-treatment and control groups, respectively. This miRNA has been linked to juvenile asthma development, as injection of let-7i-5p-loaded EVs into a juvenile asthmatic mouse model aggravated asthma symptoms [88]. EVs derived from the blood of patients with eA exhibit both DEMs and DEPs. Elevated levels of GAL10, EPO, major basic protein 1, eosinophil-derived neurotoxin, and arachidonate 15-lipoxygenase have been reported in the EVs of a patient with eA serum [61]. As mentioned above, GAL10 and EPO are over-expressed in EVs derived from the serum of patients with CRSwNP, indicating their potential as biomarkers and therapeutic targets for individuals with co-existing CRSwNP and asthma.
Several studies have explored the potential therapeutic applications of EVs for asthma. MSC-derived EVs decrease the IL-5 and IL-13 levels, inflammatory cell infiltration in bronchoalveolar lavage fluid (BALF), and ILC2 and goblet cell hyperplasia in lung epithelial cells, ultimately reducing airway hyper-responsiveness in asthmatic mouse models. These effects are associated with miR-146a-5p expression in MSC-derived EVs [77]. Furthermore, EVs derived from hypoxic MSCs, which also contain miR-146a-5p, ameliorate T2I by reducing eosinophil and TH2 mediator counts in the BALF of asthmatic mouse models [89]. As miR-146a-5p is downregulated in the EVs derived from the mucus of patients with AR [69], approaches to increase its levels may inhibit T2AI development. In addition to MSCs, ADSCs exert potent effects against asthma. ADSCs-derived EVs decrease the total IgE levels in serum and IL-4 levels in lung-draining lymph nodes and BALF and they increase Treg proportions in the lung-draining lymph nodes of asthmatic mouse models [90]. These findings underscore the potential of EV-based therapies targeting key inflammatory pathways and cellular mechanisms for asthma management.
Effects of EVs on asthma have been investigated using mast cells. Exosomes derived from bone marrow-derived mast cells were found to mitigate airway hyper-responsiveness and inflammation by binding to free IgE on exosomal FcεR in asthmatic mouse models [91]. However, EVs derived from a mouse mast cell line aggravated airway inflammation in an asthmatic mouse model by decreasing the antioxidant enzyme levels and increasing the inflammatory cell counts via the delivery of miR-21, one of the up-regulated DEMs in asthma [92]. These reports indicate the complex effects of mast cell-derived EVs on T2AI, underscoring the need for further studies to analyze their roles and therapeutic potential.

5. Conclusions

Taken together, this review describes the roles of EVs of various origins in the pathogenesis and progression of the T2AI diseases, CRSwNP, AR, and asthma. Additionally, it highlights their potential applications in the clinical diagnosis and treatment of these conditions. Figure 3 illustrates these potential diagnostic and therapeutic applications of EVs for T2AI diseases. Furthermore, this review provides a concise overview of the methods used to isolate and identify EVs in recent T2AI studies.
Many studies have used EVs of various origins, including patient-derived samples and animal models, to investigate asthma and AR. However, only a few have used EVs to study CRSwNP compared to AR and asthma. Also, verification of EV use is predominantly based on animal models or in vitro experiments. This may be attributed to challenges in the technical detection and isolation of EVs, as well as the reproducibility of EV acquisition, which is complicated by their heterogeneity. Moreover, in clinical studies including T2AI diseases, EVs isolated from patients are able to exhibit significant heterogeneity and reduced reproducibility, largely due to the diverse range of phenotypes and endotypes expressed within the patient population [93] and owing to non-standardized EV isolation and identification methods [94].
Therefore, it is crucial for further studies to focus on rigorously validating the findings from in vitro and animal models through translational medicine, ultimately advancing these results into clinical applications by establishing standard EV isolation and analysis methods and increasing the reproducibility EVs obtain [95]. Consequently, additional clinical studies will be essential to thoroughly evaluating the diagnostic and therapeutic potential of EVs, confirming their specific roles, and accelerating their integration into clinical practice.

Author Contributions

Conceptualization, J.C. and T.H.K.; formal analysis, J.C., B.K. and T.H.K.; funding acquisition, T.H.K.; investigation, J.C.; methodology, J.C. and B.K.; project administration, T.H.K.; supervision, T.H.K.; validation, J.L. and J.S.; visualization, J.C.; writing—original draft, J.C.; writing—review and editing, J.C., B.K. and T.H.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Basic Science Research Program of the National Research Foundation of Korea and funded by the Ministry of Science and Technology and the Ministry of Science, ICT and Future Planning (2017R1A2B2003575 and NRF-2020R1A2C1006398), the Ministry of Science and ICT (2020R1C1C1012288), Korea, under the ICT Creative Consilience program (IITP-2024-2020-0-01819) supervised by the Institute for Information and Communications Technology Planning and Evaluation (IITP), the Korea Health Technology R&D Project (HI17C0387 and HR22C1302), the Korea Health Industry Development Institute, and the Ministry of Health and Welfare. Additionally, this research was supported by a Korea University grant and a grant from the Korea University Medical Center and Anam Hospital in Seoul, Republic of Korea.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data used in this study are available upon reasonable request from the corresponding author. All figures included in this article were created with Biorender (BioRender.com).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Samitas, K.; Carter, A.; Kariyawasam, H.H.; Xanthou, G. Upper and Lower Airway Remodelling Mechanisms in Asthma, Allergic Rhinitis and Chronic Rhinosinusitis: The One Airway Concept Revisited. Allergy 2018, 73, 993–1002. [Google Scholar] [CrossRef] [PubMed]
  2. Fokkens, W.J.; Lund, V.J.; Hopkins, C.; Hellings, P.W.; Kern, R.; Reitsma, S.; Toppila-Salmi, S.; Bernal-Sprekelsen, M.; Alobid, J.M.I.; Anselmo-Lima Terezinha, W.; et al. European Position Paper on Rhinosinusitis and Nasal Polyps 2020. Rhinology 2020, 58 (Suppl. S29), 1–464. [Google Scholar] [CrossRef] [PubMed]
  3. Alnahas, S.; Abouammoh, N.; Althagafi, W.; Abd-Ellatif, E.E. Prevalence, Severity, and Risk Factors of Allergic Rhinitis among Schoolchildren in Saudi Arabia: A National Cross-Sectional Study, 2019. World Allergy Organ. J. 2023, 16, 100824. [Google Scholar] [CrossRef] [PubMed]
  4. Reddel, H.K.; Bacharier, L.B.; Bateman, E.D.; Brightling, C.E.; Brusselle, G.G.; Buhl, R.; Cruz, A.A.; Duijts, L.; Drazen, J.M.; FitzGerald, J.M.; et al. Global Initiative for Asthma Strategy 2021: Executive Summary and Rationale for Key Changes. Am. J. Respir. Crit. Care Med. 2022, 205, 17–35. [Google Scholar] [CrossRef]
  5. Feng, C.H.; Miller, M.D.; Simon, R.A. The United Allergic Airway: Connections between Allergic Rhinitis, Asthma, And Chronic Sinusitis. Am. J. Rhinol. Allergy 2012, 26, 187–190. [Google Scholar] [CrossRef] [PubMed]
  6. Hekking, P.P.; Amelink, M.; Wener, R.R.; Bouvy, M.L.; Bel, E.H. Comorbidities in Difficult-to-Control Asthma. J. Allergy Clin. Immunol. Pract. 2018, 6, 108–113. [Google Scholar] [CrossRef]
  7. Pinart, M.; Benet, M.; Annesi-Maesano, I.; von Berg, A.; Berdel, D.; Carlsen, K.C.L.; Carlsen, K.H.; Bindslev-Jensen, C.; Eller, E.; Fantini, M.P.; et al. Comorbidity of Eczema, Rhinitis, and Asthma in IgE-Sensitised and Non-IgE-Sensitised Children in MeDALL: A Population-Based Cohort Study. Lancet Respir. Med. 2014, 2, 131–140. [Google Scholar] [CrossRef]
  8. Heffler, E.; Pizzimenti, S.; Guida, G.; Bucca, C.; Rolla, G. Prevalence of Over-/Misdiagnosis of Asthma in Patients Referred to an Allergy Clinic. J. Asthma 2015, 52, 931–934. [Google Scholar] [CrossRef]
  9. Jaruvongvanich, V.; Mongkolpathumrat, P.; Chantaphakul, H.; Klaewsongkram, J. Extranasal Symptoms of Allergic Rhinitis Are Difficult to Treat and Affect Quality of Life. Allergol. Int. 2016, 65, 199–203. [Google Scholar] [CrossRef] [PubMed]
  10. Shehzad, A.; Islam, S.U.; Shahzad, R.; Khan, S.; Lee, Y.S. Extracellular Vesicles in Cancer Diagnostics and Therapeutics. Pharmacol. Ther. 2021, 223, 107806. [Google Scholar] [CrossRef]
  11. Czerwaty, K.; Dżaman, K.; Miechowski, W. Application of Extracellular Vesicles in Allergic Rhinitis: A Systematic Review. Int. J. Mol. Sci. 2022, 24, 367. [Google Scholar] [CrossRef] [PubMed]
  12. Kim, S.H.; Keum, B.; Kwak, S.; Byun, J.; Shin, J.M.; Kim, T.H. Therapeutic Applications of Extracellular Vesicles in Inflammatory Bowel Disease. Int. J. Mol. Sci. 2024, 25, 745. [Google Scholar] [CrossRef] [PubMed]
  13. van Niel, G.; Carter, D.R.F.; Clayton, A.; Lambert, D.W.; Raposo, G.; Vader, P. Challenges and Directions in Studying Cell–Cell Communication by Extracellular Vesicles. Nat. Rev. Mol. Cell Biol. 2022, 23, 369–382. [Google Scholar] [CrossRef] [PubMed]
  14. Carnino, J.M.; Miyawaki, S.; Rampam, S. The Role of Extracellular Vesicles in Diseases of the Ear, Nose, and Throat. Med. Sci. 2022, 11, 6. [Google Scholar] [CrossRef]
  15. Hewitt, R.J.; Lloyd, C.M. Regulation of Immune Responses by the Airway Epithelial Cell Landscape. Nat. Rev. Immunol. 2021, 21, 347–362. [Google Scholar] [CrossRef]
  16. Stanfield, C.L. Principles of Human Physiology; Pearson: San Antonio, TX, USA, 2012; ISBN 0321897714. [Google Scholar]
  17. Bousquet, J.; Anto, J.M.; Bachert, C.; Baiardini, I.; Bosnic-Anticevich, S.; Walter Canonica, G.; Melén, E.; Palomares, O.; Scadding, G.K.; Togias, A.; et al. Allergic Rhinitis. Nat. Rev. Dis. Primers 2020, 6, 95. [Google Scholar] [CrossRef] [PubMed]
  18. Holgate, S.T.; Wenzel, S.; Postma, D.S.; Weiss, S.T.; Renz, H.; Sly, P.D. Asthma. Nat. Rev. Dis. Primers 2015, 1, 15025. [Google Scholar] [CrossRef]
  19. Roan, F.; Obata-Ninomiya, K.; Ziegler, S.F. Epithelial Cell–Derived Cytokines: More than Just Signaling the Alarm. J. Clin. Investig. 2019, 129, 1441–1451. [Google Scholar] [CrossRef] [PubMed]
  20. Théry, C.; Amigorena, S. The Cell Biology of Antigen Presentation in Dendritic Cells. Curr. Opin. Immunol. 2001, 13, 45–51. [Google Scholar] [CrossRef]
  21. Kapsenberg, M.L. Dendritic-Cell Control of Pathogen-Driven T-Cell Polarization. Nat. Rev. Immunol. 2003, 3, 984–993. [Google Scholar] [CrossRef]
  22. Lambrecht, B.N.; Hammad, H. Biology of Lung Dendritic Cells at the Origin of Asthma. Immunity 2009, 31, 412–424. [Google Scholar] [CrossRef] [PubMed]
  23. Holgate, S.T. Innate and Adaptive Immune Responses in Asthma. Nat. Med. 2012, 18, 673–683. [Google Scholar] [CrossRef]
  24. Gurram, R.K.; Zhu, J. Orchestration between ILC2s and Th2 Cells in Shaping Type 2 Immune Responses. Cell. Mol. Immunol. 2019, 16, 225–235. [Google Scholar] [CrossRef] [PubMed]
  25. Kemeny, D.M. The Role of the T Follicular Helper Cells in Allergic Disease. Cell. Mol. Immunol. 2012, 9, 386–389. [Google Scholar] [CrossRef] [PubMed]
  26. Klein, U.; Dalla-Favera, R. Germinal Centres: Role in B-Cell Physiology and Malignancy. Nat. Rev. Immunol. 2008, 8, 22–33. [Google Scholar] [CrossRef]
  27. Bachert, C.; Marple, B.; Schlosser, R.J.; Hopkins, C.; Schleimer, R.P.; Lambrecht, B.N.; Bröker, B.M.; Laidlaw, T.; Song, W.J. Adult Chronic Rhinosinusitis. Nat. Rev. Dis. Primers 2020, 6, 86. [Google Scholar] [CrossRef]
  28. Simon, H.U.; Yousefi, S.; Schranz, C.; Schapowal, A.; Bachert, C.; Blaser, K. Direct Demonstration of Delayed Eosinophil Apoptosis as a Mechanism Causing Tissue Eosinophilia. J. Immunol. 1997, 158, 3902–3908. [Google Scholar] [CrossRef] [PubMed]
  29. Takafuji, S.; Ohtoshi, T.; Takizawa, H.; Tadokoro, K.; Ito, K. Eosinophil Degranulation in the Presence of Bronchial Epithelial Cells. Effect of Cytokines and Role of Adhesion. J. Immunol. 1996, 156, 3980–3985. [Google Scholar] [CrossRef]
  30. Modena, B.D.; Dazy, K.; White, A.A. Emerging Concepts: Mast Cell Involvement in Allergic Diseases. Transl. Res. 2016, 174, 98–121. [Google Scholar] [CrossRef] [PubMed]
  31. Karpman, D.; Ståhl, A.L.; Arvidsson, I. Extracellular Vesicles in Renal Disease. Nat. Rev. Nephrol. 2017, 13, 545–562. [Google Scholar] [CrossRef] [PubMed]
  32. Cheng, L.; Hill, A.F. Therapeutically Harnessing Extracellular Vesicles. Nat. Rev. Drug Discov. 2022, 21, 379–399. [Google Scholar] [CrossRef] [PubMed]
  33. He, F.; Li, L.; Fan, R.; Wang, X.; Chen, X.; Xu, Y. Extracellular Vesicles: An Emerging Regenerative Treatment for Oral Disease. Front. Cell Dev. Biol. 2021, 9, 669011. [Google Scholar] [CrossRef] [PubMed]
  34. Gilmore, W.J.; Bitto, N.J.; Kaparakis-Liaskos, M. Apoptotic Bodies: Mechanism of Formation, Isolation and Functional Relevance. In New Frontiers: Extracellular Vesicles; Springer: Cham, Switzerland, 2021; Volume 97, ISBN 9783030671709. [Google Scholar]
  35. Piccin, A.; Murphy, W.G.; Smith, O.P. Circulating Microparticles: Pathophysiology and Clinical Implications. Blood Rev. 2007, 21, 157–171. [Google Scholar] [CrossRef] [PubMed]
  36. Huttner, W.B.; Zimmerberg, J. Implications of Lipid Microdomains for Membrane Curvature, Budding and Fission. Curr. Opin. Cell Biol. 2001, 13, 478–484. [Google Scholar] [CrossRef] [PubMed]
  37. Wollert, T.; Hurley, J.H. Molecular Mechanism of Multivesicular Body Biogenesis by ESCRT Complexes. Nature 2010, 464, 864–869. [Google Scholar] [CrossRef] [PubMed]
  38. Larios, J.; Mercier, V.; Roux, A.; Gruenberg, J. ALIX- and ESCRT-III-Dependent Sorting of Tetraspanins to Exosomes. J. Cell Biol. 2020, 219, e201904113. [Google Scholar] [CrossRef] [PubMed]
  39. Verderio, C.; Gabrielli, M.; Giussani, P. Role of Sphingolipids in the Biogenesis and Biological Activity of Extracellular Vesicles. J. Lipid Res. 2018, 59, 1325–1340. [Google Scholar] [CrossRef] [PubMed]
  40. Andreu, Z.; Yáñez-Mó, M. Tetraspanins in Extracellular Vesicle Formation and Function. Front. Immunol. 2014, 5, 442. [Google Scholar] [CrossRef]
  41. Johnstone, R.M.; Adam, M.; Hammond, J.R.; Orr, L.; Turbide, C. Vesicle Formation during Reticulocyte Maturation. Association of Plasma Membrane Activities with Released Vesicles (Exosomes). J. Biol. Chem. 1987, 262, 9412–9420. [Google Scholar] [CrossRef]
  42. Ng, S.C.; Ananthakrishnan, A.N. New approaches along the IBD course: Diet, tight control and stem cells. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 82–84. [Google Scholar] [CrossRef]
  43. Tancini, B.; Buratta, S.; Sagini, K.; Costanzi, E.; Delo, F.; Urbanelli, L.; Emiliani, C. Insight into the Role of Extracellular Vesicles in Lysosomal Storage Disorders. Genes 2019, 10, 510. [Google Scholar] [CrossRef] [PubMed]
  44. Alexander, M.; Hu, R.; Runtsch, M.C.; Kagele, D.A.; Mosbruger, T.L.; Tolmachova, T.; Seabra, M.C.; Round, J.L.; Ward, D.M.; O’Connell, R.M. Exosome-delivered microRNAs modulate the inflammatory response to endotoxin. Nat. Commun. 2015, 6, 7321. [Google Scholar] [CrossRef] [PubMed]
  45. Théry, C.; Witwer, K.W.; Aikawa, E.; Alcaraz, M.J.; Anderson, J.D.; Andriantsitohaina, R.; Antoniou, A.; Arab, T.; Archer, F.; Atkin-Smith, G.K.; et al. Minimal Information for Studies of Extracellular Vesicles 2018 (MISEV2018): A Position Statement of the International Society for Extracellular Vesicles and Update of the MISEV2014 Guidelines. J. Extracell. Vesicles 2018, 7, 1535750. [Google Scholar] [CrossRef] [PubMed]
  46. Qian, K.; Fu, W.; Li, T.; Zhao, J.; Lei, C.; Hu, S. The roles of small extracellular vesicles in cancer and immune regulation and translational potential in cancer therapy. J. Exp. Clin. Cancer Res. 2022, 41, 286. [Google Scholar] [CrossRef]
  47. Buzas, E.I. The roles of extracellular vesicles in the immune system. Nat. Rev. Immunol. 2023, 23, 236–250. [Google Scholar] [CrossRef]
  48. Schleimer, R.P. Immunopathogenesis of Chronic Rhinosinusitis and Nasal Polyposis. Annu. Rev. Pathol. 2017, 12, 331–357. [Google Scholar] [CrossRef]
  49. Promsopa, C.; Kansara, S.; Citardi, M.J.; Fakhri, S.; Porter, P.; Luong, A. Prevalence of Confirmed Asthma Varies in Chronic Rhinosinusitis Subtypes. Int. Forum Allergy Rhinol. 2016, 6, 373–377. [Google Scholar] [CrossRef]
  50. Kato, A.; Schleimer, R.P.; Bleier, B.S. Mechanisms and Pathogenesis of Chronic Rhinosinusitis. J. Allergy Clin. Immunol. 2022, 149, 1491–1503. [Google Scholar] [CrossRef]
  51. Nakayama, T.; Yoshikawa, M.; Asaka, D.; Okushi, T.; Matsuwaki, Y.; Otori, N.; Hama, T.; Moriyama, H. Mucosal Eosinophilia and Recurrence of Nasal Polyps—New Classification of Chronic Rhinosinusitis. Rhinology 2011, 49, 392–396. [Google Scholar] [CrossRef]
  52. Stevens, W.W.; Peters, A.T.; Tan, B.K.; Klingler, A.I.; Poposki, J.A.; Hulse, K.E.; Grammer, L.C.; Welch, K.C.; Smith, S.S.; Conley, D.B.; et al. Associations between Inflammatory Endotypes and Clinical Presentations in Chronic Rhinosinusitis. J. Allergy Clin. Immunol. Pract. 2019, 7, 2812–2820.e3. [Google Scholar] [CrossRef]
  53. Lin, Y.T.; Lin, C.F.; Liao, C.K.; Yeh, T.H. Comprehensive Evaluation of Type 2 Endotype and Clinical Features in Patients with Chronic Rhinosinusitis with Nasal Polyps in Taiwan: A Cross-Sectional Study. Eur. Arch. Otorhinolaryngol. 2023, 280, 5379–5389. [Google Scholar] [CrossRef] [PubMed]
  54. Miyake, M.M.; Workman, A.D.; Nocera, A.L.; Wu, D.; Mueller, S.K.; Finn, K.; Amiji, M.M.; Bleier, B.S. Discriminant Analysis Followed by Unsupervised Cluster Analysis Including Exosomal Cystatins Predict Presence of Chronic Rhinosinusitis, Phenotype, and Disease Severity. Int. Forum Allergy Rhinol. 2019, 9, 1069–1076. [Google Scholar] [CrossRef]
  55. Kato, Y.; Takabayashi, T.; Sakashita, M.; Imoto, Y.; Tokunaga, T.; Ninomiya, T.; Morikawa, T.; Yoshida, K.; Noguchi, E.; Fujieda, S. Expression and Functional Analysis of CST1 in Intractable Nasal Polyps. Am. J. Respir. Cell Mol. Biol. 2018, 59, 448–457. [Google Scholar] [CrossRef]
  56. Kato, T.; Imatani, T.; Minaguchi, K.; Saitoh, E.; Okuda, K. Salivary Cystatins Induce Interleukin-6 Expression via Cell Surface Molecules in Human Gingival Fibroblasts. Mol. Immunol. 2002, 39, 423–430. [Google Scholar] [CrossRef] [PubMed]
  57. Mueller, S.K.; Nocera, A.L.; Dillon, S.T.; Libermann, T.A.; Wendler, O.; Bleier, B.S. Tissue and Exosomal Serine Protease Inhibitors Are Significantly Overexpressed in Chronic Rhinosinusitis with Nasal Polyps. Am. J. Rhinol. Allergy 2019, 33, 359–368. [Google Scholar] [CrossRef]
  58. Mueller, S.K.; Nocera, A.L.; Workman, A.; Libermann, T.; Dillon, S.T.; Stegmann, A.; Wurm, J.; Iro, H.; Wendler, O.; Bleier, B.S. Significant Polyomic and Functional Upregulation of the PAPP-A/IGFBP-4/5/IGF-1 Axis in Chronic Rhinosinusitis with Nasal Polyps. Int. Forum Allergy Rhinol. 2020, 10, 546–555. [Google Scholar] [CrossRef]
  59. He, S.; Wu, J.; Han, D.; Li, Y.; Wang, T.; Wei, H.; Pan, Y.; Zang, H. Differential Expression Profile of Plasma Exosomal MicroRNAs in Chronic Rhinosinusitis with Nasal Polyps. Exp. Biol. Med. 2022, 247, 1039–1046. [Google Scholar] [CrossRef] [PubMed]
  60. Shi, C.Y.; Kingston, E.R.; Kleaveland, B.; Lin, D.H.; Stubna, M.W.; Bartel, D.P. The ZSWIM8 Ubiquitin Ligase Mediates Target-Directed MicroRNA Degradation. Science 2020, 370, eabc9359. [Google Scholar] [CrossRef] [PubMed]
  61. Yoshimura, H.; Takeda, Y.; Shirai, Y.; Yamamoto, M.; Nakatsubo, D.; Amiya, S.; Enomoto, T.; Hara, R.; Adachi, Y.; Edahiro, R.; et al. Galectin-10 in Serum Extracellular Vesicles Reflects Asthma Pathophysiology. J. Allergy Clin. Immunol. 2024, 153, 1268–1281. [Google Scholar] [CrossRef] [PubMed]
  62. Liu, F.T.; Stowell, S.R. The Role of Galectins in Immunity and Infection. Nat. Rev. Immunol. 2023, 23, 479–494. [Google Scholar] [CrossRef]
  63. Ye, L.; Wang, H.; Li, H.; Liu, H.; Lv, T.; Song, Y.; Zhang, F. Eosinophil Peroxidase Over-Expression Predicts the Clinical Outcome of Patients with Primary Lung Adenocarcinoma. J. Cancer 2019, 10, 1032–1038. [Google Scholar] [CrossRef] [PubMed]
  64. Zhou, M.; Tan, K.S.; Guan, W.J.; Jiang, L.J.; Deng, J.; Gao, W.X.; Lee, Y.M.; Xu, Z.F.; Luo, X.; Liu, C.; et al. Proteomics Profiling of Epithelium-Derived Exosomes from Nasal Polyps Revealed Signaling Functions Affecting Cellular Proliferation. Respir. Med. 2020, 162, 105871. [Google Scholar] [CrossRef]
  65. Bousquet, P.J.; Leynaert, B.; Neukirch, F.; Sunyer, J.; Janson, C.M.; Anto, J.; Jarvis, D.; Burney, P. Geographical Distribution of Atopic Rhinitis in the European Community Respiratory Health Survey I. Allergy 2008, 63, 1301–1309. [Google Scholar] [CrossRef]
  66. Vandenplas, O.; Vinnikov, D.; Blanc, P.D.; Agache, I.; Bachert, C.; Bewick, M.; Cardell, L.O.; Cullinan, P.; Demoly, P.; Descatha, A.; et al. Impact of Rhinitis on Work Productivity: A Systematic Review. J. Allergy Clin. Immunol. Pract. 2018, 6, 1274–1286.e9. [Google Scholar] [CrossRef]
  67. Devillier, P.; Bousquet, J.; Salvator, H.; Naline, E.; Grassin-Delyle, S.; de Beaumont, O. In Allergic Rhinitis, Work, Classroom and Activity Impairments Are Weakly Related to Other Outcome Measures. Clin. Exp. Allergy 2016, 46, 1456–1464. [Google Scholar] [CrossRef] [PubMed]
  68. Wu, G.; Yang, G.; Zhang, R.; Xu, G.; Zhang, L.; Wen, W.; Lu, J.; Liu, J.; Yu, Y. Altered MicroRNA Expression Profiles of Extracellular Vesicles in Nasal Mucus from Patients with Allergic Rhinitis. Allergy Asthma Immunol. Res. 2015, 7, 449–457. [Google Scholar] [CrossRef] [PubMed]
  69. Li, H.; Wang, Y.; Han, X. ESP-B4 Promotes Nasal Epithelial Cell-Derived Extracellular Vesicles Containing MiR-146a-5p to Modulate Smad3/GATA-3 Thus Relieving Allergic Rhinitis: ESP-B4/MiR-146a-5p in AR. Phytomedicine 2023, 108, 154516. [Google Scholar] [CrossRef] [PubMed]
  70. Zhu, X.; Wang, X.; Wang, Y.; Zhao, Y. Exosomal Long Non-coding RNA GAS5 Suppresses Th1 Differentiation and Promotes Th2 Differentiation via Downregulating EZH2 and T-Bet in Allergic Rhinitis. Mol. Immunol. 2020, 118, 30–39. [Google Scholar] [CrossRef] [PubMed]
  71. Mattick, J.S.; Amaral, P.P.; Carninci, P.; Carpenter, S.; Chang, H.Y.; Chen, L.L.; Chen, R.; Dean, C.; Dinger, M.E.; Fitzgerald, K.A.; et al. Long Non-coding RNAs: Definitions, Functions, Challenges and Recommendations. Nat. Rev. Mol. Cell Biol. 2023, 24, 430–447. [Google Scholar] [CrossRef]
  72. Mowel, W.K.; Kotzin, J.J.; McCright, S.J.; Neal, V.D.; Henao-Mejia, J. Control of Immune Cell Homeostasis and Function by LncRNAs. Trends Immunol. 2018, 39, 55–69. [Google Scholar] [CrossRef]
  73. Miossec, P.; Van den Berg, W.T. Th1/Th2 Cytokine Balance in Arthritis. Arthritis Rheum. 1997, 40, 2105–2115. [Google Scholar] [CrossRef]
  74. Busse, W.W.; Lemanske, R.F.J.R. Asthma. N. Engl. J. Med. 2001, 344, 350–362. [Google Scholar] [CrossRef]
  75. Fang, S.B.; Zhou, Z.R.; Peng, Y.Q.; Liu, X.Q.; He, B.X.; Chen, D.H.; Chen, D.; Fu, Q.L. Plasma EVs Display Antigen-Presenting Characteristics in Patients with Allergic Rhinitis and Promote Differentiation of Th2 Cells. Front. Immunol. 2021, 12, 710372. [Google Scholar] [CrossRef]
  76. Ogi, K.; Ramezanpour, M.; Liu, S.; Ferdoush Tuli, J.; Bennett, C.; Suzuki, M.; Fujieda, S.; Psaltis, A.J.; Wormald, P.J.; Vreugde, S. Der p 1 Disrupts the Epithelial Barrier and Induces IL-6 Production in Patients With House Dust Mite Allergic Rhinitis. Front. Allergy 2021, 2, 692049. [Google Scholar] [CrossRef]
  77. Fang, S.B.; Zhang, H.Y.; Wang, C.; He, B.X.; Liu, X.Q.; Meng, X.C.; Peng, Y.Q.; Xu, Z.B.; Fan, X.L.; Wu, Z.J.; et al. Small Extracellular Vesicles Derived from Human Mesenchymal Stromal Cells Prevent Group 2 Innate Lymphoid Cell-Dominant Allergic Airway Inflammation through Delivery of MiR-146a-5p. J. Extracell. Vesicles 2020, 9, 1723260. [Google Scholar] [CrossRef] [PubMed]
  78. Lyu, B.; Wei, Z.; Jiang, L.; Ma, C.; Yang, G.; Han, S. MicroRNA-146a negatively regulates IL-33 in activated group 2 innate lymphoid cells by inhibiting IRAK1 and TRAF6. Genes Immun. 2022, 21, 37–44. [Google Scholar] [CrossRef] [PubMed]
  79. Peng, Y.Q.; Wu, Z.C.; Xu, Z.B.; Fang, S.B.; Chen, D.H.; Zhang, H.Y.; Liu, X.Q.; He, B.X.; Chen, D.; Akdis, C.A.; et al. Mesenchymal Stromal Cells-Derived Small Extracellular Vesicles Modulate DC Function to Suppress Th2 Responses via IL-10 in Patients with Allergic Rhinitis. Eur. J. Immunol. 2022, 52, 1129–1140. [Google Scholar] [CrossRef]
  80. Chapoval, S.; Dasgupta, P.; Dorsey, N.J.; Keegan, A.D. Regulation of the T Helper Cell Type 2 (Th2)/T Regulatory Cell (Treg) Balance by IL-4 and STAT6. J. Leukoc. Biol. 2010, 87, 1011–1018. [Google Scholar] [CrossRef] [PubMed]
  81. Wu, J.; Huang, Q.M.; Liu, Y.; Zhou, J.; Tang, W.R.; Wang, X.Y.; Wang, L.F.; Zhang, Z.H.; Tan, H.L.; Guan, X.H.; et al. Long-Term Hypoxic HUCMSCs-Derived Extracellular Vesicles Alleviates Allergic Rhinitis through Triggering Immunotolerance of Their VEGF-Mediated Inhibition of Dendritic Cells Maturation. Int. Immunopharmacol. 2023, 124, 110875. [Google Scholar] [CrossRef]
  82. Yang, W.; Pan, Z.; Zhang, J.; Wang, L.; Lai, J.; Zhou, S.; Zhang, Z.; Fan, K.; Deng, D.; Gao, Z.; et al. Extracellular Vesicles from Adipose Stem Cells Ameliorate Allergic Rhinitis in Mice by Immunomodulatory. Front. Immunol. 2023, 14, 1302336. [Google Scholar] [CrossRef]
  83. Stern, J.; Pier, J.; Litonjua, A.A. Asthma Epidemiology and Risk Factors. Semin. Immunopathol. 2020, 42, 5–15. [Google Scholar] [CrossRef]
  84. Oppenheimer, J.; Hoyte, F.C.L.; Phipatanakul, W.; Silver, J.; Howarth, P.; Lugogo, N.L. Allergic and Eosinophilic Asthma in the Era of Biomarkers and Biologics: Similarities, Differences and Misconceptions. Ann. Allergy Asthma Immunol. 2022, 129, 169–180. [Google Scholar] [CrossRef]
  85. Kulshreshtha, A.; Ahmad, T.; Agrawal, A.; Ghosh, B. Proinflammatory Role of Epithelial Cell-Derived Exosomes in Allergic Airway Inflammation. J. Allergy Clin. Immunol. 2013, 131, 1194–1203.e14. [Google Scholar] [CrossRef] [PubMed]
  86. Bartel, S.; La Grutta, S.; Cilluffo, G.; Perconti, G.; Bongiovanni, A.; Giallongo, A.; Behrends, J.; Kruppa, J.; Hermann, S.; Chiang, D.; et al. Human Airway Epithelial Extracellular Vesicle MiRNA Signature Is Altered upon Asthma Development. Allergy 2020, 75, 346–356. [Google Scholar] [CrossRef]
  87. Schindler, V.E.M.; Alhamdan, F.; Preußer, C.; Hintz, L.; Alashkar Alhamwe, B.; Nist, A.; Stiewe, T.; Pogge von Strandmann, E.; Potaczek, D.P.; Thölken, C.; et al. Side-Directed Release of Differential Extracellular Vesicle-Associated MicroRNA Profiles from Bronchial Epithelial Cells of Healthy and Asthmatic Subjects. Biomedicines 2022, 10, 622. [Google Scholar] [CrossRef]
  88. Zheng, R.; Du, M.; Tian, M.; Zhu, Z.; Wei, C.; Chu, H.; Gan, C.; Liang, J.; Xue, R.; Gao, F.; et al. Fine Particulate Matter Induces Childhood Asthma Attacks via Extracellular Vesicle-Packaged Let-7i-5p-Mediated Modulation of the MAPK Signaling Pathway. Adv. Sci. 2022, 9, e2102460. [Google Scholar] [CrossRef] [PubMed]
  89. Dong, L.; Wang, Y.; Zheng, T.; Pu, Y.; Ma, Y.; Qi, X.; Zhang, W.; Xue, F.; Shan, Z.; Liu, J.; et al. Hypoxic HUCMSC-Derived Extracellular Vesicles Attenuate Allergic Airway Inflammation and Airway Remodeling in Chronic Asthma Mice. Stem Cell Res. Ther. 2021, 12, 4. [Google Scholar] [CrossRef] [PubMed]
  90. Mun, S.J.; Kang, S.A.; Park, H.K.; Yu, H.S.; Cho, K.S.; Roh, H.J. Intranasally Administered Extracellular Vesicles from Adipose Stem Cells Have Immunomodulatory Effects in a Mouse Model of Asthma. Stem Cells Int. 2021, 2021, 6686625. [Google Scholar] [CrossRef] [PubMed]
  91. Xie, G.; Yang, H.; Peng, X.; Lin, L.; Wang, J.; Lin, K.; Cui, Z.; Li, J.; Xiao, H.; Liang, Y.; et al. Mast Cell Exosomes Can Suppress Allergic Reactions by Binding to IgE. J. Allergy Clin. Immunol. 2018, 141, 788–791. [Google Scholar] [CrossRef]
  92. Zou, Y.; Zhou, Q.; Zhang, Y. MicroRNA-21 Released from Mast Cells-Derived Extracellular Vesicles Drives Asthma in Mice by Potentiating Airway Inflammation and Oxidative Stress. Am. J. Transl. Res. 2021, 13, 7475–7491. [Google Scholar] [PubMed]
  93. Sangaphunchai, P.; Todd, I.; Fairclough, L.C. Extracellular vesicles and asthma: A review of the literature. Clin. Exp. Allergy 2020, 50, 291–307. [Google Scholar] [CrossRef] [PubMed]
  94. Rezaie, J.; Feghhi, M.; Etemadi, T. A review on exosomes application in clinical trials: Perspective, questions, and challenges. Cell Commun. Signal. 2022, 20, 145. [Google Scholar] [CrossRef] [PubMed]
  95. Lee, K.W.A.; Chan, L.K.W.; Hung, L.C.; Phoebe, L.K.W.; Park, Y.; Yi, K.-H. Clinical Applications of Exosomes: A Critical Review. Int. J. Mol. Sci. 2024, 25, 7794. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The pathogenesis of type 2 airway inflammation (T2AI). Various triggers, including pollens, viruses, bacteria, antigens/allergens, smoke, house dust mites (HDMs), particulate matter (PM), and stress, disrupt the airway epithelium, leading to the release of chemokines, such as the C-C motif chemokine ligand (CCL)-27, and alarmin cytokines, such as thymic stromal lymphopoietin (TSLP), interleukin (IL)-25, and IL-33. These signals activate type 2 innate lymphoid cells (ILC2s) and attract dendritic cells (DCs) expressing chemokine receptors, such as the C-C motif chemokine receptor (CCR)-10, to the site of disruption. Immature DCs (imDCs) migrate to the disrupted epithelium and are converted to mature DCs (mDCs) via stimulation by alarmin cytokines and interactions with the infected antigens/allergens. These mDCs present the antigens to naïve T-cells in local lymph nodes, promoting their polarization into TH2 cells. These TH2 cells, along with ILC2s, secrete cytokines (IL-4, IL-5, and IL-13), which influence various cell types, such as epithelial cells and eosinophils. Additionally, IL-4 and IL-13 facilitate B-cell class switching to IgE. Plasma cells secrete IgE, which binds to the IgE receptors (FcεR) on mast cell membranes via a process known as sensitization. Upon re-exposure to the antigen or allergen, the cross-linking of IgE on mast cells induces their degranulation, releasing inflammatory mediators, such as histamines (Hs), prostaglandins (PGs), and leukotrienes (LTs). Additionally, IL-5 activates and degranulates eosinophils. This integrative degranulation of mast cells and eosinophils ultimately leads to inflammatory symptoms in the airway. HDM, house dust mite; PM, particulate matter; CCL, C-C motif chemokine ligand; TSLP, thymic stromal lymphopoietin; IL, interleukin; ILC2, type 2 innate lymphoid cell; DC, dendritic cell; CCR, C-C motif chemokine receptor; imDC, immature DC; mDC, mature DC; Ig, immunoglobulin; FcεR, IgE receptor; H, histamine; PG, prostaglandin; LT, leukotriene.
Figure 1. The pathogenesis of type 2 airway inflammation (T2AI). Various triggers, including pollens, viruses, bacteria, antigens/allergens, smoke, house dust mites (HDMs), particulate matter (PM), and stress, disrupt the airway epithelium, leading to the release of chemokines, such as the C-C motif chemokine ligand (CCL)-27, and alarmin cytokines, such as thymic stromal lymphopoietin (TSLP), interleukin (IL)-25, and IL-33. These signals activate type 2 innate lymphoid cells (ILC2s) and attract dendritic cells (DCs) expressing chemokine receptors, such as the C-C motif chemokine receptor (CCR)-10, to the site of disruption. Immature DCs (imDCs) migrate to the disrupted epithelium and are converted to mature DCs (mDCs) via stimulation by alarmin cytokines and interactions with the infected antigens/allergens. These mDCs present the antigens to naïve T-cells in local lymph nodes, promoting their polarization into TH2 cells. These TH2 cells, along with ILC2s, secrete cytokines (IL-4, IL-5, and IL-13), which influence various cell types, such as epithelial cells and eosinophils. Additionally, IL-4 and IL-13 facilitate B-cell class switching to IgE. Plasma cells secrete IgE, which binds to the IgE receptors (FcεR) on mast cell membranes via a process known as sensitization. Upon re-exposure to the antigen or allergen, the cross-linking of IgE on mast cells induces their degranulation, releasing inflammatory mediators, such as histamines (Hs), prostaglandins (PGs), and leukotrienes (LTs). Additionally, IL-5 activates and degranulates eosinophils. This integrative degranulation of mast cells and eosinophils ultimately leads to inflammatory symptoms in the airway. HDM, house dust mite; PM, particulate matter; CCL, C-C motif chemokine ligand; TSLP, thymic stromal lymphopoietin; IL, interleukin; ILC2, type 2 innate lymphoid cell; DC, dendritic cell; CCR, C-C motif chemokine receptor; imDC, immature DC; mDC, mature DC; Ig, immunoglobulin; FcεR, IgE receptor; H, histamine; PG, prostaglandin; LT, leukotriene.
Ijms 25 09455 g001
Figure 2. The biogenesis and Classification of extracellular vesicles (EVs). EVs are classified into apoptotic bodies (ApoBDs), micro-vesicles, and exosomes based on their origin and size. ApoBDs are primarily formed via the shedding and blebbing of apoptotic cells, whereas micro-vesicles are produced via cellular budding. Exosomes are typically formed and released through the endosomal sorting complex required for the transport (ESCRT)-dependent pathway, with markers, such as tetraspanins, the tumor susceptibility gene 101 (TSG101), and the ALG-2-interacting protein X (ALIX), playing key roles in their identification. Recently, EV classification has become more nuanced, taking into account their size, density, biochemical composition, and condition of parent cells. EV, extracellular vesicle; ApoBD; apoptotic body; ESCRT, endosomal sorting complex required for transport; TSG, tumor susceptibility gene; ALIX, ALG-2-interacting protein X; lEV, large EV; mEV, medium EV; sEV, small EV.
Figure 2. The biogenesis and Classification of extracellular vesicles (EVs). EVs are classified into apoptotic bodies (ApoBDs), micro-vesicles, and exosomes based on their origin and size. ApoBDs are primarily formed via the shedding and blebbing of apoptotic cells, whereas micro-vesicles are produced via cellular budding. Exosomes are typically formed and released through the endosomal sorting complex required for the transport (ESCRT)-dependent pathway, with markers, such as tetraspanins, the tumor susceptibility gene 101 (TSG101), and the ALG-2-interacting protein X (ALIX), playing key roles in their identification. Recently, EV classification has become more nuanced, taking into account their size, density, biochemical composition, and condition of parent cells. EV, extracellular vesicle; ApoBD; apoptotic body; ESCRT, endosomal sorting complex required for transport; TSG, tumor susceptibility gene; ALIX, ALG-2-interacting protein X; lEV, large EV; mEV, medium EV; sEV, small EV.
Ijms 25 09455 g002
Figure 3. Comprehensive applications of EVs for various T2AI diseases. EVs exhibit differentially expressed molecules in chronic rhinosinusitis with nasal polyps (CRSwNP), allergic rhinitis (AR), and asthma and can be applied for the diagnosis of these conditions. Specifically, mesenchymal stromal cells (MSCs) and adipose tissue-derived stem cells (ADSCs)-derived EVs are potential therapeutic targets for AR and asthma. ↑: Up-regulation, ↓: Down-regulation.
Figure 3. Comprehensive applications of EVs for various T2AI diseases. EVs exhibit differentially expressed molecules in chronic rhinosinusitis with nasal polyps (CRSwNP), allergic rhinitis (AR), and asthma and can be applied for the diagnosis of these conditions. Specifically, mesenchymal stromal cells (MSCs) and adipose tissue-derived stem cells (ADSCs)-derived EVs are potential therapeutic targets for AR and asthma. ↑: Up-regulation, ↓: Down-regulation.
Ijms 25 09455 g003
Table 3. List of studies investigating EVs in asthma.
Table 3. List of studies investigating EVs in asthma.
No.DonorUsed EVsEV Isolation
Methods
EV Identification MethodsSubjectsResults
Effects
Ref.
1Healthy hBECs (BEAS-2B)Exosomes/sEVsCentrifugation UltracentrifugationBead-based assay (HSP70, CD63)
WB (CD63, TSG101, ALIX, RAB5B, RAB27A)
IL-13 treatment after vs. beforeExosome secretion ↑[85]
2Healthy primary hBECsExosomes/sEVsBasal/apical medium with ExoQuick-TC kitNTA (5–1000 nm),
TEM,
WB (CD63, CD9, ALIX),
SEC (10 kDa, 0.2 µm)
IL-13 treatment after vs. beforemiR-34a ↓ in basal EVs
miR-92b ↓ in basal EVs
miR-210 ↓ in apical EVs
[86]
3Primary hBECsExosomes/sEVsBasal/apical medium centrifugation
(500× g for 10 min, 2000× g for 20 min, 10,000× g for 30 min)
Filtration with 0.22-µm filter,
ExoEasy kit
NanoFCM (50–250 nm),
Bead-based flow cytometry
(CD9, CD63, CD81)
Patients with asthma vs. ControlmiR-9 ↑ in apical EVs
let-7, miR-9, and miR-10 ↑ in basal EVs
[87]
4Healthy hBECs
plasma
Exosomes/sEVsCulture of hBECs with ExoQuick TC Kit ExoQuick Plasma Prep with Thrombin KitTEM, NanoFCM,
WB
(CD63, TSG101, ALIX)
PM2.5 treatment after vs. before
Patients with asthma vs. Control
let-7i-5p ↑[88]
5SerumExosomes/sEVsPhosphatidylserine (PS) affinity methodTEM (CD9)
NTA (<200 nm)
Patients with eA vs. ControlGAL10, EPO, MBP1, EDN, ALOX15 ↑
in EVs
[61]
6hMSCssEVsCentrifugation
(2000× g for 20 min)
Anion-exchange chromatography
Ultracentrifugation
(300× g for 5 min, 2000× g for 20 min, 12,000× g for 30 min, 110,000× g for 70 min)
Flow cytometry
(CD9, CD63, CD81),
ELISA (CD63),
WB (CD9, CD63, CD81, ALIX, TSG101) Protein concentration, NTA, TEM
EV treatment after vs. before
in asthmatic mice
miR-146a-5p ↑ in MSC-derived EVs
Infiltration of inflammatory cells ↓ in peritracheal area and BALF
IL-5 and IL-13 ↓ in BALF
Epithelial goblet cells ↓ in lungs
ILC2 and airway hyperresponsiveness ↓ in lungs
[77]
7Hypoxic-hMSCsUnknownUltracentrifugation
(300× g for 10 min, 2000× g for 20 min, 100,000× g for 90 min)
TEM,
WB (HSP70 TSG101),
NTA (150–160 nm)
Hypoxic MSC vs. Normal MSC
EV treatment after vs. before
in asthmatic mice
miR-146a-5p ↑ in hypoxic MSC-EVs
Total cells, eosinophils, and TH2 mediator ↓ in BALF
Airway inflammation ↓
[89]
8Murine ADSCsUnknownFiltration with
0.45 and 0.22-µm filters
Ultracentrifugation (100,000× g for 2 h)
TEM (100–400 nm), NTA (127 nm),
WB (CD40 and CD81)
EV treatment after vs. before
in asthmatic mice
Total IgE and IgG ↓ in serum
Inflammatory cells, eosinophils, and IL-4 ↓ in BALF Eosinophilic Lung inflammation ↓
IL-4 ↓ lung draining lymph nodes
Treg ↑ in lung draining lymph nodes
[90]
9Murine BMMCsExosomes/sEVsCentrifugation TEM (50–80 nm),
WB (FcεRI),
Flow cytometry,
Confocal microscopy
Exosome treatment after vs. before
in asthmatic mice
Airway inflammation ↓
Airway hyperresponsiveness ↓
[91]
10Murine mast cell (MC/9)UnknownCentrifugation
(300× g for 10 min, 20,000× g for 20 min)
Filtration with a 0.2-µm filter
Ultracentrifugation (100,000× g for 90 min)
TEM, NTA, WB
(HSP70, CD9, TSG101)
EV treatment after vs. before
in mouse epithelial cell line/asthmatic mice
MiR-21 ↑ in mouse mast cell-derived EVs/asthmatic Mice
miR-21 ↑ after MC-EV treatment in mouse epithelial cells
Antioxidant enzymes ↓/inflammatory cells ↑ in asthmatic mice
[92]
A vs. B: Relative results of group A vs. group B. Abbreviation: No., Number; EVs, extracellular vesicles; hBECs, human bronchial epithelial cells; hMSCs, human mesenchymal stromal cells; ADSCs, adipose tissue-derived stem cells; BMMCs, bone marrow-derived mast cells; sEVs, small EVs; NTA, nanoparticle tracking analysis; TEM, transmission electron microscopy; SEC, size exclusion chromatography; WB, western blotting; FACS, fluorescence-activated cell sorting; CD, cluster of differentiation; ALIX, ALG-2 interacting protein X; TSG, tumor susceptibility gene; HSP, heat-shock protein; RAB, Ras-related protein; IL, interleukin; PM2.5, particulate matter less than 2.5 µm in diameter; eA, eosinophilic asthma, miR, microRNA; GAL10, galectin 10; EPO, eosinophil peroxidase; MBP1, major basic protein 1; EDN, eosinophil-derived neurotoxin; ALOX15, arachidonate 15-lipoxygenase; BALF, bronchoalveolar lavage fluid; Ig, immunoglobulin; Treg, regulatory T-cell; Ref., reference. ↑: Up-regulation, ↓: Down-regulation.
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

Cheon, J.; Kim, B.; Lee, J.; Shin, J.; Kim, T.H. Functions and Clinical Applications of Extracellular Vesicles in TH2 Cell-Mediated Airway Inflammatory Diseases: A Review. Int. J. Mol. Sci. 2024, 25, 9455. https://doi.org/10.3390/ijms25179455

AMA Style

Cheon J, Kim B, Lee J, Shin J, Kim TH. Functions and Clinical Applications of Extracellular Vesicles in TH2 Cell-Mediated Airway Inflammatory Diseases: A Review. International Journal of Molecular Sciences. 2024; 25(17):9455. https://doi.org/10.3390/ijms25179455

Chicago/Turabian Style

Cheon, Jaehwan, Byoungjae Kim, Juhyun Lee, Jaemin Shin, and Tae Hoon Kim. 2024. "Functions and Clinical Applications of Extracellular Vesicles in TH2 Cell-Mediated Airway Inflammatory Diseases: A Review" International Journal of Molecular Sciences 25, no. 17: 9455. https://doi.org/10.3390/ijms25179455

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