Skip to main content

Immunological factors, important players in the development of asthma

Abstract

Asthma is a heterogeneous disease, and its development is the result of a combination of factors, including genetic factors, environmental factors, immune dysfunction and other factors. Its specific mechanism has not yet been fully investigated. With the improvement of disease models, research on the pathogenesis of asthma has made great progress. Immunological disorders play an important role in asthma. Previously, we thought that asthma was mainly caused by an imbalance between Th1 and Th2 immune responses, but this theory cannot fully explain the pathogenesis of asthma. Recent studies have shown that T-cell subsets such as Th1 cells, Th2 cells, Th17 cells, Tregs and their cytokines contribute to asthma through different mechanisms. For the purpose of the present study, asthma was classified into distinct phenotypes based on airway inflammatory cells, such as eosinophilic asthma, characterized by predominant eosinophil aggregates, and neutrophilic asthma, characterized by predominant neutrophil aggregates. This paper will examine the immune mechanisms underlying different types of asthma, and will utilize data from animal models and clinical studies targeting specific immune pathways to inform more precise treatments for this condition.

Peer Review reports

Introduction

Asthma is a common respiratory ailment that affects more than 300 million people globally. This ailment manifests itself through chronic airway inflammation, airway hyperresponsiveness (AHR), reversible airflow limitation, etc. Notably, the syndrome manifests as the infiltration of diverse inflammatory cells, including eosinophils, neutrophils, and lymphocytes, and the involvement of inflammatory mediators and cytokines. Asthma has a significant impact on human health worldwide, and there is an urgent need to improve our understanding of its pathogenesis. The establishment of asthma phenotypes and endotypes provides potential for increased accuracy in disease classification and the exploration of suitable biomarkers and therapeutic targets [1,2,3]. T-cell subtypes, including Th1 cells, Th2 cells, Th17 cells, Tregs, and their corresponding cytokines, are significant contributors to the development of asthma. Drug therapy targeting immune molecules offers new therapeutic prospects for asthma patients [4,5,6].

Imbalance in the immune response of Th1 and Th2 cells

T helper type 1 (Th1) cells and Th2 cells are important effector cells that inhibit each other during the development of asthma [7]. Th1 cells mainly mediate the cellular immune response, delay hypersensitivity and assist in the synthesis of phagocytosis-related antibodies, e.g., IgG, IgM and IgA, which may inhibit the development of asthma [8]. Th2 cells primarily mediate the humoral immune response and immediate hypersensitivity [9]. They also aid in the proliferation of B cells, IgE synthesis, antibody production and non-phagocytic host defense, contributing to the development of asthma. These cells are derived from the same group of precursor cells (Th0) and are induced by different cytokines, and the direction of T-cell polarization is closely related to the local cytokine environment in which dendritic cells (DCs) play a significant role. According to previous research, type I dendritic cells (DCIs) exhibit poor maturation and insufficient IL-12 secretion, which prevents naïve CD4 + T cells (Th0) from differentiating into Th1 cells [10]. However, when Th0 cells are stimulated with IL-4, DCII promotes the differentiation of Th0 cells into Th2 cells, which causes Th1 cells to secrete less IFN-γ and Th2 cells to secrete more IL-4 [11, 12]. Several additional factors have been shown to affect the selective development of Th1 and Th2 cells. These factors include the dosage and affinity of the antigen, major histocompatibility complex (MHC) haplotypes, and co-stimulatory factors. The Th1 cell-specific transcription factor T-bet binds to the Th2 cell-specific transcription factor GATA-3, hindering the binding of GATA-3 to Th2 cytokine genes and suppressing the production of Th2 cytokines, and vice versa. A decrease in the level of Th1 cells expressing T-bet in the airways of asthma patients caused an increase in GATA-3 expression and Th2 cytokine production. Increased GATA-3 inhibits Th1 cytokine production through the inhibition of STAT4, resulting in a further increase in respiratory Th2 cells in asthma patients [13].

For many years, an imbalance in Th1/Th2 cells has been acknowledged to be a significant contributor to the onset of asthma. Nevertheless, further research has indicated that the Th1/Th2 cell imbalance theory is unable to account for various experimental phenomena. In addition, the pathogenesis of asthma has progressed to include attention given to Th17 cells and Tregs.

The main mechanism of eosinophilic asthma is the type 2 inflammatory response mediated by Th2 cells and ILC2s

Eosinophilic asthma is the most common type of asthma found in patients with asthma of varying severity and is now thought to be primarily associated with a type 2 inflammatory response mediated by Th2 cells and Type II innate lymphocytes (ILC2s) [14,15,16].

Foreign antigens enter the body and are captured by antigen-presenting cells (APCs), such as DCs. Here, they form MHC II-antigen-peptide complexes with the help of MHC class II molecules. T cell receptors (TCRs) on naïve CD4 + T cells (Th0) recognize these complexes. In the presence of signals generated by co-stimulatory molecules (CD80, CD86, OX40 and its ligands) and in the presence of IL-4, Th2 cells and follicular helper T (Tfh) cells differentiate. Consequently, they release type 2 cytokines, which trigger the maturation and recruitment of other immune cells involved in the allergic cascade. Additionally, these cytokines promote the accumulation of activated Th2 cells, eosinophils, ILC2s, DCs, macrophages and natural killer T cells (NKTs) in lung tissue. Frequent tissue damage triggers a reparative course that accelerates the deposition of connective tissue and the modification of airway configuration, resulting in thickening of the basement membrane and smooth muscle hypertrophy, recognized as airway remodeling [17, 18].

Allergen-specific Th2 cells, which produce IL-4, IL-5, IL-13 and even IL-9, have been found in the blood of individuals with allergic asthma [19]. The function of these cytokines in asthma has been extensively debated. Studies have indicated that IL-4 primarily originates from mast cells, NKT cells, γδT cells, Th2 cells, ILC2s and basophils. It interacts with B cells, induces IgE class switch recombination, B-cell maturation, and promotes the expression of adhesion molecules (intercellular adhesion molecule-1 [ICAM-1] and vascular cell adhesion molecule-1 [VCAM-1]) to facilitate eosinophil extravasation [20]. Both IL-4 and IL-13 bind to the IL-4Rα chain to initiate the corresponding response. This finding suggests that there may be partial overlap in their functions. IL-13 plays an important role in increased mucus production, goblet cell metaplasia and airway hyperresponsiveness. Furthermore, periostin gene expression in bronchial epithelial cells is upregulated by IL-13 and IL-4 [21]. Periostin acts on fibroblasts to promote airway remodeling, enhance mucus secretion and recruit eosinophils. IL-5 facilitates eosinophil development, activation, migration, proliferation and survival. Research conducted in mice has shown that IL-5 promotes eosinophil recruitment and activation, but its role in promoting AHR has been controversial [22]. The contribution of Th9 cells, a distinctive subset of T-cells that produce IL-9, to asthma pathogenesis remains debatable as there is a paucity of research studies covering this topic. When anti-IL-9 antibodies were administered, exposure to allergens in mouse models resulted in positive outcomes, such as decreased expression of pertinent cytokines, decreased lung eosinophil counts, and decreased AHR [23]. However, a study indicated that the forced expiratory volume in 1 s (FEV1) level, incidence of asthma exacerbation, and quality of life in patients failed to benefit from treatment with the humanized IL-9 neutralizing antibody MEDI-528, also known as enoclizumab [24]. These studies suggest a potential correlation between IL-9 and AHR, pulmonary eosinophilia, and mucus hypersecretion.

Tfh cells are a distinct subpopulation of CD4 + T cells that orchestrate paracrine signaling from B cells to stimulate plasma cell differentiation and IgE production. In a mouse model of house dust mite (HDM) sensitization, Tfh cells were shown to reside in the lungs draining mediastinal lymph nodes subsequent to the initial sensitization phase. These cells then differentiate into Th2 cells and migrate to the lungs to execute their functions following a second exposure to HDM allergens [25]. It is noteworthy that Tfh cells induce IgE production, even in patients where conventional Th2 cells are not present or have limited effector cell capacity [26].

ILC2s represent a recently identified lymphocyte subset producing type 2 cytokines, particularly IL-5 and IL-13. Several studies have shown that ILC2s play an important role in the development of asthma. When airway epithelial cells are harmed by allergens, pollutants, or infections, pattern recognition receptors (PRRs) on the cells can identify pathogen-associated molecular patterns (PAMPs) present in inhaled bacteria and parasites, as well as damage-associated molecular patterns (DAMPs) produced by dead or injured cells. The activation of downstream signals and the release of cytokines, including IL-25, IL-33 and thymic stromal lymphopoietin (TSLP), by epithelial cells stimulate the differentiation and activation of ILC2s and other innate immune cells (such as macrophages, basophils, and eosinophils). This leads to the enhancement of the type 2 immune response [27,28,29,30][Fig. 1].

TSLP, IL-25, and IL-33 are regarded as primary molecules involved in type 2 inflammatory response and have a crucial role in the pathogenesis of asthma. IL-25 levels were found to be higher in individuals with eosinophilic asthma. Elevated IL-25 levels are associated with enhanced airway hyperresponsiveness, increased numbers of airway and blood eosinophils, elevated serum IgE levels, and subepithelial stromal deposits [31]. IL-25 acts directly on bronchial epithelial cells (BECs) and circulating fibroblasts (CFs) in autocrine and paracrine-dependent manners, respectively, promoting airway remodeling and fibrosis in patients with asthma [32]. IL-33 is elevated in the epithelial cells and bronchoalveolar lavage (BAL) fluid of individuals with asthma and is clearly correlated with the severity of the disease [33]. As a resident in lung tissue, ILC2s express significantly high levels of IL-33R. Activation of ERK1/2, p38 MAPK, AKT, JNK, and nuclear factor-κB (NF-κB) by IL-33 can cause chemotaxis of ILC2s, ultimately promoting the accumulation of immunological cells in the lung [34]. TSLP mRNA expression increases in the airway mucosa of asthmatic individuals and is positively correlated with disease severity [35]. Studies have shown that TSLP promotes ILC2 lifespan and glucocorticoid resistance [36]. Furthermore, although TSLP has traditionally been linked to type 2 inflammation, its expression in BAL fluid has been found to also have an association with neutrophil inflammation [37].

ILC2s not only mediate eosinophil aggregation, goblet cell metaplasia, and airway hyperresponsiveness through the production of IL-5, IL-9, and IL-13 but also may act as APCs to provide CD4 + T lymphocytes with MHCII molecules and OX40 ligands (OX40L). Activation of the ILC2s in the lung has been reported to induce key asthma features such as eosinophil aggregation and airway hyperresponsiveness in mouse models lacking T cells and B cells [17]. Therefore, we believe that the roles of Th2 cells and ILC2s in the pathogenesis of asthma are complementary.

Fig. 1
figure 1

Involvement of Th2 and ILC2 cell-associated cytokines in the development of asthma

Drugs that target type 2 cytokines and their receptors are now integral to asthma treatment. A humanized monoclonal IL-13 neutralizing antibody called lebrikizumab was found to reduce asthma exacerbations and improve lung function in patients with moderate-to-severe asthma in a clinical trial. Lebrikizumab also demonstrated good therapeutic effects in asthma patients with high levels of periosteal proteins [38]. Lebrikizumab treatment had no impact on tissue or blood eosinophil counts, which could account for its ineffectiveness in managing asthma exacerbations. In contrast, dupilumab, a monoclonal antibody that targets the IL-4R alpha chain common to both the IL-4 and IL-13 receptors, prevents transduction signals stimulated by IL-4 and IL-13, and has been utilized for asthma treatment. However, it did not have any significant impact on those with non-eosinophilic asthma. In eosinophilic asthma patients, the drug lessened the number of asthma attacks and glucocorticoid usage. Mepolizumab, a human IL-5 monoclonal antibody, significantly decreases blood eosinophil counts, decreases the incidence of acute exacerbations, and enhances the quality of life of patients with severe eosinophilic asthma [39, 40].

Clinical trials of drugs targeting type 2 cytokines and their receptors have yielded useful results, and a variety of drugs targeting key molecules of the type 2 inflammatory response are now in use, providing better options for the treatment of asthma, which in turn confirms the role of these cytokines in the pathogenesis of asthma. In addition to therapeutic methods focused on type 2 cytokines, which are the molecules that trigger type 2 inflammatory responses, investigations concerning epithelial-derived cytokines such as IL-25, IL-33, and TSLP have been conducted and have provided insight into novel drug development prospects [41, 42].

The role of Th17 cells and Th1 cells in neutrophilic asthma

Some asthma patients do not exhibit tissue eosinophilia, but instead exhibit neutrophilia. This type of asthma may be associated with severe asthma and glucocorticoid-resistant asthma. The development of this part of asthma may be related to the immune response of Th17 cells and Th1 cells, and the exact mechanism still needs to be further investigated [43, 44].

The mechanism by which Th17 cells recruit neutrophils is not fully understood. IL-17A protein expression in sputum from asthmatics was found to be positively correlated with increased neutrophil count and airway hyperresponsiveness [45]. Overexpression of IL-17A in a mouse model resulted in the generation of a large number of peripheral neutrophils and enhanced granulopoiesis, suggesting that IL-17A significantly promoted neutrophil maturation, migration, and proliferation [46]. Increased airway neutrophil infiltration may result from the ability of IL-17A to increase the secretion of neutrophil chemokines, such as CXCL-8, and from the ability of airway epithelial cells, IL-1β, IL-6, and GM-CSF to proliferate. Reactive oxygen species (ROS), neutrophil elastase (NE), and matrix metalloproteinase 9 (MMP-9) are produced by neutrophils [47, 48]. MMP-9 stimulates DC migration and maturation, and NE causes mucus secretion by goblet cells. These processes negatively regulate tissue inhibitor of matrix metalloproteinases 1 (TIMP-1), which in turn promotes mucus production, bronchoconstriction, and tissue remodeling. Furthermore, neutrophil extracellular traps (NETs)—structures made of DNA, modified histones, and granule proteins such as NE and myeloperoxidase (MPO)—can be produced by activated neutrophils. These structures can disturb the integrity of bronchial epithelial cells, leading to the release of cytokines such as TSLP and IL-33, which can exacerbate the pathology of asthma [49]. Moreover, discoveries from a mouse model indicate that IL-17A intensifies smooth muscle contraction and worsens airway hyperreactivity by activating NF-κB, which results in an increase of RhoA and ROCK2 expression [50].

However, whether the cytokines IL-17F and IL-22 produced by Th17 cells are protective or pathogenic in asthma pathogenesis remains to be further investigated. IL-17F inhibits the activation of DCs, which results in a diminished Th2 response and decreased airway inflammation. These findings may provide a possible explanation for the increased allergic responses found in IL-17F-deficient mice in some studies [51]. However, it has been shown that IL-17F enhances airway smooth muscle contraction, migration and proliferation, thereby promoting asthmatic features such as airway hyperresponsiveness and airway remodeling. In a mouse model of sensitization to ovalbumin (OVA), IL-22 contributes significantly to the inflammation of airways triggered by allergen inhalation through eosinophils and neutrophils [52]. However, a different study found that introducing the IL-22 gene before systemic sensitization restrained antigen-stimulated CD4 + T cell expansion and the inflammation in the airways caused by eosinophils [53].

However, the function of Th17 cells in the pathogenesis of asthma remains controversial. While in models of OVA antigen sensitization, Th17 cells encourage neutrophil recruitment to the airways and contribute to the development of AHR, in other models, the Th17 response has a negative regulating effect on the Th2 response, thus halting AHR development. In clinical trials conducted on neutrophilic asthmatic patients, the administration of a CXCR2 antagonist to block CXCL8-mediated neutrophil recruitment did not affect any asthma outcomes. Therefore, the role of neutrophils in asthma remains uncertain, and it is unknown whether they are active participants or passive bystanders [54].

On the other hand, TNF-α and IFN-γ represent crucial cytokines that are discharged during the.

Fig. 2
figure 2

Th1 and Th17 promote neutrophilic inflammation

Th1 immune response and cause the release of pro-inflammatory mediators, airway hyperresponsiveness and airway remodeling [55]. Patients suffering from neutrophilic asthma exhibit elevated sputum TNF-α levels, and it is believed that TNF-α is involved in neutrophilic asthma by promoting concurrent neutrophil recruitment with the cytokine IL-17A. It has been demonstrated that IFN-γ + CD4 + T cells are predominant in the BAL fluid of children with severe asthma suffering from both eosinophilic and neutrophilic asthma [Fig. 2]. This study suggests that the Th1 immune response is related to severe asthma [56]. Furthermore, TNF-α and IFN-γ levels are increased in certain patients who are resistant to corticosteroid therapy. The investigation presents the possibility that interactions between TNF-α and IFN-γ play a role in corticosteroid resistance among asthma patients [57]. It has been shown that the Th1 cytokine IFN-γ promotes neutrophil recruitment in the presence of the cytokine IL-17 [58].

However, the exact potential of blocking IL-17A or its receptor for treating neutrophilic asthma remains unclear [59]. Whether and how neutrophilic inflammation can be a target for asthma treatment requires further in-depth study. In addition, studying the relationship between neutrophilic asthma and hormone resistance may be useful in the treatment of refractory asthma.

The role of regulatory T cells (Tregs) in the development of asthma

Tregs are a specialized population of CD4 + T cells that act to suppress the immune response, thereby maintaining the body’s immune homeostasis and self-tolerance. Reduced numbers of Tregs have been observed in the blood and induced sputum of patients with asthma, and during acute asthma exacerbations, asthmatics have a greater degree of Tregs deficiency [60, 61]. In addition to their reduced numbers, the ability of Tregs to suppress effector T cells and their cytokines is significantly reduced in allergic individuals [62, 63]. In a mouse model, adoptive transfer of CD4 + CD25 + Tregs to mice prior to antigen challenge inhibited the development of allergic disease, and adoptive transfer of Tregs after disease onset attenuated established inflammation and airway remodeling [64, 65]. These studies support the idea that decreased numbers of Tregs and functional deficits contribute to asthma. Tregs can be divided into two main subpopulations: thymus-derived Tregs (tTregs) and peripherally-derived Tregs (pTregs). tTregs primarily recognize self-antigens, which is important for maintaining self-tolerance and preventing autoimmunity. However, pTregs are thought to be involved primarily in the tolerogenic response to foreign antigens and microbes [66, 67].

Tregs may mediate immunomodulatory functions through a variety of mechanisms, including the release of inhibitory cytokines (IL-10, TGF-β, and IL-35, etc.), the release of cytolytic molecules [granzymes (Gzm) A and B], and the regulation of the expression of specific transcription factors and receptors and so on. Tregs can modulate the activation of various cell types involved in allergic responses, such as eosinophils, neutrophils, mast cells, and B cells, through the mechanisms mentioned above [68].

IL-10, TGF-β, and IL-35 are considered the major factors involved in mediating the regulatory functions of Tregs [69].

IL-10 is a cytokine produced by various cell types and is implicated in the regulatory functions of Tregs. The transfer of allergen-specific CD4 + CD25 + Tregs into OVA-sensitized mice significantly reduced AHR and inflammatory cell recruitment, and IL-10 levels were elevated in the BAL fluid of the transferred mice. Anti-IL-10R treatment abolished the inhibition of AHR and lung eosinophil infiltration by CD4 + CD25 + regulatory cells when neutralizing antibodies against the IL-10 receptor were administered during the allergen challenge phase [64]. Tregs release of IL-10 obstructs the production of pro-inflammatory cytokines and restricts the antigen-presenting function of APCs. Furthermore, IL-10 regulates the function of diverse immune cells, including mast cells, Th2 cells, eosinophils, and DCs, which are implicated in allergic responses [70, 71].

Transforming growth factor-β (TGF-β) is a pleiotropic cytokine released by Tregs that is involved in the regulation of the immune response. On the one hand, it directly prevents the proliferation of T cells and B cells and inhibits the proliferation and function of macrophages by reducing the release of various cytokines. In addition, TGF-β is involved in the regulation of airway inflammation, tissue repair and other processes. However, the contribution of TGF-β to the immunosuppressive function of Tregs remains controversial. It has been reported that the lack of TGF-β in mice does not affect their immunosuppressive capacity [72]. On the other hand, TGF-β not only has anti-asthmatic effects in the pathogenesis of asthma, but also has pro-asthmatic effects. TGF-β is an important participant in airway remodeling. The current study shows a significant correlation between the number of epithelial or submucosal cells expressing TGF-β, the thickness of the basement membrane and the number of fibroblasts in asthma [73]. TGF-β induces an increase in the number of fibroblasts and airway smooth muscle proliferation, which in turn leads to extracellular matrix (ECM) deposition and ultimately exacerbates the process of airway remodeling [74].

IL-35 is a vital cytokine that Tregs release and directly affects effector cells. In patients with allergic asthma, the protein and mRNA levels of IL-35 are lower than those in healthy controls [75]. Additionally, it has been recognized that this cytokine is produced mainly by Tregs and has anti-inflammatory and immunosuppressive effects. Research has suggested that the absence of one of the two subunits diminishes the inhibitory ability of Tregs, affirming the function of IL-35 in immune system suppression in an inflammatory bowel disease (IBD) mouse model [76]. IL-10 and TGF-β are not essential for inhibiting IL-17-dependent AHR after airway sensitization; in particular, IL-35 performs as the primary cytokine for this purpose [77]. Therefore, these studies demonstrate the crucial involvement of IL-35 in the pathogenesis of asthma.

Stimulation by various cytokines results in the differentiation of naïve T cells into discrete subpopulations under the guidance of lineage-specific transcription factors, including T-bet (Th1 cells), GATA3 (Th2 cells), RORγt (Th17 cells) and FOXP3 (Tregs). The differentiation mechanisms induced by transcription factors are mutually exclusive, resulting in functional suppression between T-cell subsets. This phenomenon is especially noticeable between effector Th cells and Tregs, for instance, between Th2 cells and Tregs and between Th17 cells and Tregs. Imbalances in the immune response of these effector and regulatory cells have been linked to the development of asthma.

Both Th17 cells and Tregs require TGF-β for differentiation, but this process depends on the presence or absence of IL-6 [78]. It has been shown that the presence of IL-6 promotes the differentiation of primitive T cells to Th17 cells. However, in an environment lacking IL-6, primitive T cells differentiate into Tregs.

In addition, the Th2 cytokine IL-4 first activates the transcription factor STAT6, followed by the activation of GATA-3, which leads to the differentiation of primitive T cells into Th2 cells and promotes the expression of Th2 cell-related genes. TGF-β can bind to its receptor, activate Smad family transcription factors, promote FOXP3 transcription, and promote Treg-related gene expression. Moreover, GATA-3 can directly bind to the FOXP3 promoter and inhibit the expression of Treg-related genes, thus inhibiting the differentiation of primitive T cells into Tregs. FOXP3 also binds to GATA-3 and prevents it from regulating the expression of Th2 genes. Since GATA-3 and FOXP3 inhibit each other’s functional activity, the relative amounts of GATA-3 and FOXP3 may determine the fate of CD4 + T cells to differentiate into Th2 cells or Tregs [79].

The balance between the immune response of Th cells and the immunosuppression of Tregs is important for the maintenance of immune homeostasis. However, in recent years, it has been discovered that transcription factors produced by effector Th cells and their corresponding cytokines are involved not only in the immune responses of effector Th cells but also in the immunosuppression of Tregs to prevent excessive immune responses [80]. The treatment of asthma by targeting Tregs is currently immature. Rapamycin is a novel macrolide immunosuppressive drug that has been widely used for the prevention of clinical allograft rejection and some autoimmune diseases. Experiments in mouse models have shown that rapamycin promotes Tregs differentiation and induces organismal immune tolerance. Further studies on the mechanism of Tregs regulation may provide new ideas for clinical treatment [81].

The role of additional factors in the development of asthma

Eosinophils

Eosinophils play a pivotal role in asthma, as their accumulation in lung tissue is stimulated by IL-5, thereby fostering eosinophil formation in the bone marrow and attracting eosinophils to the lung mucosa and mesenchyme via eosinophil chemokines (e.g., eotaxins). Upon stimulation, eosinophils release large amounts of inflammatory mediators, including cytokines (such as IL-13 and IL-5), chemokines, granule mediators, and cysteinyl leukotrienes (cysLTs). Eosinophils produce cytotoxic proteins, including major basic protein (MBP), eosinophil peroxidase (EPX), eosinophil cationic protein (ECP) and eosinophil-derived neurotoxin (EDN), which are held in intracellular granules. These proteins contribute to the remodeling of airways, increased airway responsiveness and mucus generation [82]. Eosinophils also release extracellular deposits of genetic material that interweave to form a network known as eosinophil extracellular traps (EETs). EETs have an inherent function in fostering eosinophil degranulation and prompting the production of IL-6 and IL-8 by epithelial cells. It not only has an important function in immunity against extracellular pathogens but also contributes to the pathogenesis of asthma [83]. Increased production of EETs can increase mucus viscosity, leading to airway obstruction. In patients with severe asthma, the number of peripheral eosinophils producing EETs is elevated, and these cells stimulate lung epithelial cells to generate IL-33 and TSLP [84]. In addition, eosinophils can impact the performance of other immune cells through cytokines and chemokines. For instance, they can promote Th2 cell differentiation by secreting the Th2-associated cytokine IL-4.

Mast cells and basophils

Mast cells and basophils play comparable roles in asthma pathogenesis and participate in the organism’s reaction to allergens. During the sensitization phase, IgE binds to the IgE high-affinity receptor FcεRI, which both mast cells and basophils express. Following re-exposure to allergens, cross-linking between antigenic peptides and allergen-specific IgE molecules provokes the degranulation of mast cells and basophils. Pre-synthesized mediators found in mast cells are released, consisting of histamine and proteases. Additionally, newly formed mediators, including cytokines, chemokines, prostaglandins, and leukotrienes, are unleashed, serving to further stimulate the inflammatory response characteristic of asthma. The release of mast cell mediators induces a local inflammatory response, accompanied by smooth muscle contraction, amplified capillary dilation and permeability, and heightened glandular secretions. These responses augment the infiltration of lymphocytes and eosinophils into the inflamed site [85]. Mast cell degranulation is directly related to asthma severity. The severity of asthma is closely linked to the degranulation of mast cells. Mast cell-produced chymotrypsin is a protease that undermines the integrity of the epithelial barrier, worsening EC destruction [86]. In the airway, the release of LTC4, PGD2, IL-4, and IL-13 from mast cells promotes the secretion of large amounts of mucus by goblet cells [87]. In addition, mast cells and eosinophils promote type 2 cytokine production. Mast cell localization within the airway smooth muscle is a crucial characteristic in the development of asthma. It contributes to hypertrophy and hyperplasia of smooth muscle and facilitates the establishment of AHR [88]. Basophils enhance B-cell proliferation, category-switching reorganization, plasma cell differentiation and maturation, and immunoglobulin production [89].

IgE

Elevated serum IgE levels are a crucial indicator for patient assessment and strongly correlate with asthma development [90]. Research suggests that heightened IgE levels are associated with airway hyperresponsiveness, bronchial wall thickening, and severe asthma [91, 92]. IgE has been shown to directly affect eosinophil function by activating and releasing eosinophil peroxidase, expressing integrins, and releasing TNF-α. Furthermore, IgE stimulates airway smooth muscle directly, resulting in the production of cytokines (IL-4, IL-5, IL-13, TNF-α and TSLP) and chemokines (CCL5, CCL11, CXCL8 and CXCL10). These substances are responsible for the contraction and proliferation of airway smooth muscle, eventually leading to airway remodeling [93].

Omalizumab is currently the most advanced anti-IgE monoclonal antibody of human origin. It functions by intercepting the allergic cascade process via the prevention of the binding of IgE to the FcεRI receptor on mast cells, basophils, and other inflammatory cells. Several clinical trials conducted in both children and adults have reported positive outcomes. Therefore, omalizumab is extensively used in clinical practice. Omalizumab has favorable effects, including decreasing sputum eosinophil counts, reducing airway hyperresponsiveness, and improving symptoms. Moreover, the use of omalizumab has led to a decrease in the frequency of acute exacerbations and hospitalizations, along with a reduction in steroid use and an improvement in asthma control [94].

Discussion

The development of asthma is driven by the participation of various immune cells and molecules. Despite the increased options available for investigating disease pathogenesis in current disease models, our comprehension of its pathogenesis remains inadequate. Molecular biology and immunology advancements have given way to selective blocking agents that offer an original perspective for treating asthma patients. Inhibiting crucial molecules involved in asthma pathogenesis has significantly contributed to advancements in asthma treatment. Studying the cellular and molecular mechanisms involved in asthma-related inflammation might unveil novel therapeutic targets that could enhance the efficacy of asthma drugs. Additionally, these findings could inform the quest for more suitable biomarkers.

Data availability

No datasets were generated or analysed during the current study.

References

  1. Anderson GP. Endotyping asthma: new insights into key pathogenic mechanisms in a complex, heterogeneous disease. Lancet. 2008;372(9643):1107–19. https://doi.org/10.1016/s0140-6736(08)61452-x.

    Article  PubMed  Google Scholar 

  2. Bhakta NR, Woodruff PG. Human asthma phenotypes: from the clinic, to cytokines, and back again. Immunol Rev. 2011;242(1):220–32. https://doi.org/10.1111/j.1600-065X.2011.01032.x.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Ricciardolo FLM, Guida G, Bertolini F, Di Stefano A, Carriero V. Phenotype overlap in the natural history of asthma. Eur Respir Rev. 2023;32. https://doi.org/10.1183/16000617.0201-2022.

  4. Kuruvilla ME, Lee FE, Lee GB. Understanding asthma phenotypes, endotypes, and mechanisms of Disease. Clin Rev Allergy Immunol. 2019;56(2):219–33. https://doi.org/10.1007/s12016-018-8712-1.

    Article  PubMed  PubMed Central  Google Scholar 

  5. Ji T, Li H. T-helper cells and their cytokines in pathogenesis and treatment of asthma. Front Immunol. 2023;14:1149203. https://doi.org/10.3389/fimmu.2023.1149203.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Habib N, Pasha MA, Tang DD. (2022) Current Understanding of Asthma Pathogenesis and Biomarkers. Cells.11(17).https://doi.org/10.3390/cells11172764.

  7. Busse WW, Rosenwasser LJ. Mechanisms of asthma. J Allergy Clin Immunol. 2003;111(3 Suppl). https://doi.org/10.1067/mai.2003.158.

  8. Bagheri N, Salimzadeh L, Shirzad H. The role of T helper 1-cell response in Helicobacter pylori-infection. Microb Pathog. 2018;123:1–8. https://doi.org/10.1016/j.micpath.2018.06.033.

    Article  CAS  PubMed  Google Scholar 

  9. Haase P, Voehringer D. Regulation of the humoral type 2 immune response against allergens and helminths. Eur J Immunol. 2021;51(2):273–9. https://doi.org/10.1002/eji.202048864.

    Article  CAS  PubMed  Google Scholar 

  10. Jongbloed SL, Kassianos AJ, McDonald KJ, et al. Human CD141+ (BDCA-3) + dendritic cells (DCs) represent a unique myeloid DC subset that cross-presents necrotic cell antigens. J Exp Med. 2010;207(6):1247–60. https://doi.org/10.1084/jem.20092140.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Pilette C, Jacobson MR, Ratajczak C, et al. Aberrant dendritic cell function conditions Th2-cell polarization in allergic rhinitis. Allergy. 2013;68(3):312–21. https://doi.org/10.1111/all.12090.

    Article  CAS  PubMed  Google Scholar 

  12. Kumamoto Y, Linehan M, Weinstein JS, Laidlaw BJ, Craft JE, Iwasaki A. CD301b⁺ dermal dendritic cells drive T helper 2 cell-mediated immunity. Immunity. 2013;39(4):733–43. https://doi.org/10.1016/j.immuni.2013.08.029.

    Article  CAS  PubMed  Google Scholar 

  13. Usui T, Nishikomori R, Kitani A, Strober W. GATA-3 suppresses Th1 development by downregulation of Stat4 and not through effects on IL-12Rbeta2 chain or T-bet. Immunity. 2003;18(3):415–28. https://doi.org/10.1016/s1074-7613(03)00057-8.

    Article  CAS  PubMed  Google Scholar 

  14. Komlósi ZI, van de Veen W, Kovács N, et al. Cellular and molecular mechanisms of allergic asthma. Mol Aspects Med. 2022. https://doi.org/10.1016/j.mam.2021.100995.

    Article  PubMed  Google Scholar 

  15. Agache I, Eguiluz-Gracia I, Cojanu C, et al. Advances and highlights in asthma in 2021. Allergy. 2021;76(11):3390–407. https://doi.org/10.1111/all.15054.

    Article  PubMed  Google Scholar 

  16. Boonpiyathad T, Sözener ZC, Satitsuksanoa P, Akdis CA. Immunologic mechanisms in asthma. Semin Immunol. 2019. https://doi.org/10.1016/j.smim.2019.101333.

    Article  PubMed  Google Scholar 

  17. Hammad H, Lambrecht BN. The basic immunology of asthma. Cell. 2021;184(6):1469–85. https://doi.org/10.1016/j.cell.2021.02.016.

  18. Lambrecht BN, Hammad H, Fahy JV. The cytokines of Asthma. Immunity. 2019;50(4):975–91. https://doi.org/10.1016/j.immuni.2019.03.018.

  19. Seumois G, Ramírez-Suástegui C, Schmiedel BJ, et al. Single-cell transcriptomic analysis of allergen-specific T cells in allergy and asthma. Sci Immunol. 2020;5. https://doi.org/10.1126/sciimmunol.aba6087.

  20. Godar M, Deswarte K, Vergote K, et al. A bispecific antibody strategy to target multiple type 2 cytokines in asthma. J Allergy Clin Immunol. 2018;142(4):1185–93. https://doi.org/10.1016/j.jaci.2018.06.002.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Lopez-Guisa JM, Powers C, File D, Cochrane E, Jimenez N, Debley JS. Airway epithelial cells from asthmatic children differentially express proremodeling factors. J Allergy Clin Immunol. 2012;129(4):990–7. https://doi.org/10.1016/j.jaci.2011.11.035.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Corry DB, Folkesson HG, Warnock ML, et al. Interleukin 4, but not interleukin 5 or eosinophils, is required in a murine model of acute airway hyperreactivity. J Exp Med. 1996;183(1):109–17. https://doi.org/10.1084/jem.183.1.109.

    Article  CAS  PubMed  Google Scholar 

  23. Cheng G, Arima M, Honda K, et al. Anti-interleukin-9 antibody treatment inhibits airway inflammation and hyperreactivity in mouse asthma model. Am J Respir Crit Care Med. 2002;166(3):409–16. https://doi.org/10.1164/rccm.2105079.

    Article  PubMed  Google Scholar 

  24. Oh CK, Leigh R, McLaurin KK, Kim K, Hultquist M, Molfino NA. A randomized, controlled trial to evaluate the effect of an anti-interleukin-9 monoclonal antibody in adults with uncontrolled asthma. Respir Res. 2013;14(1):93. https://doi.org/10.1186/1465-9921-14-93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Ballesteros-Tato A, Randall TD, Lund FE, Spolski R, Leonard WJ, León B. T follicular helper cell plasticity shapes pathogenic T helper 2 cell-mediated immunity to Inhaled House Dust Mite. Immunity. 2016;44(2):259–73. https://doi.org/10.1016/j.immuni.2015.11.017.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Kobayashi T, Iijima K, Dent AL, Kita H. Follicular helper T cells mediate IgE antibody response to airborne allergens. J Allergy Clin Immunol. 2017;139(1):300–13. https://doi.org/10.1016/j.jaci.2016.04.021.

    Article  CAS  PubMed  Google Scholar 

  27. Whetstone CE, Ranjbar M, Omer H, Cusack RP, Gauvreau GM. (2022) The Role of Airway Epithelial Cell Alarmins in Asthma. Cells.11(7).https://doi.org/10.3390/cells11071105.

  28. Hong H, Liao S, Chen F, Yang Q, Wang DY. Role of IL-25, IL-33, and TSLP in triggering united airway diseases toward type 2 inflammation. Allergy. 2020;75(11):2794–804. https://doi.org/10.1111/all.14526.

    Article  CAS  PubMed  Google Scholar 

  29. Heijink IH, Kuchibhotla VNS, Roffel MP, et al. Epithelial cell dysfunction, a major driver of asthma development. Allergy. 2020;75(8):1902–17. https://doi.org/10.1111/all.14421.

    Article  PubMed  Google Scholar 

  30. Duchesne M, Okoye I, Lacy P. Epithelial cell alarmin cytokines: Frontline mediators of the asthma inflammatory response. Front Immunol. 2022;13:975914. https://doi.org/10.3389/fimmu.2022.975914.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Cheng D, Xue Z, Yi L, et al. Epithelial interleukin-25 is a key mediator in Th2-high, corticosteroid-responsive asthma. Am J Respir Crit Care Med. 2014;190(6):639–48. https://doi.org/10.1164/rccm.201403-0505OC.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Yao X, Chen Q, Wang X, Liu X, Zhang L. IL-25 induces airway remodeling in asthma by orchestrating the phenotypic changes of epithelial cell and fibrocyte. Respir Res. 2023;24(1):212. https://doi.org/10.1186/s12931-023-02509-z.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Li Y, Wang W, Lv Z, et al. Elevated expression of IL-33 and TSLP in the airways of human asthmatics in vivo: a potential biomarker of severe refractory disease. J Immunol. 2018;200(7):2253–62. https://doi.org/10.4049/jimmunol.1701455.

    Article  CAS  PubMed  Google Scholar 

  34. Li Y, Chen S, Chi Y, et al. Kinetics of the accumulation of group 2 innate lymphoid cells in IL-33-induced and IL-25-induced murine models of asthma: a potential role for the chemokine CXCL16. Cell Mol Immunol. 2019;16(1):75–86. https://doi.org/10.1038/s41423-018-0182-0.

    Article  CAS  PubMed  Google Scholar 

  35. Ying S, O’Connor B, Ratoff J, et al. Thymic stromal lymphopoietin expression is increased in asthmatic airways and correlates with expression of Th2-attracting chemokines and disease severity. J Immunol. 2005;174(12):8183–90. https://doi.org/10.4049/jimmunol.174.12.8183.

    Article  CAS  PubMed  Google Scholar 

  36. Kabata H, Moro K, Fukunaga K, et al. Thymic stromal lymphopoietin induces corticosteroid resistance in natural helper cells during airway inflammation. Nat Commun. 2013. https://doi.org/10.1038/ncomms3675.

    Article  PubMed  Google Scholar 

  37. Mitchell PD, Salter BM, Oliveria JP, et al. IL-33 and its receptor ST2 after inhaled Allergen Challenge in allergic asthmatics. Int Arch Allergy Immunol. 2018;176(2):133–42. https://doi.org/10.1159/000488015.

    Article  CAS  PubMed  Google Scholar 

  38. Hanania NA, Noonan M, Corren J, et al. Lebrikizumab in moderate-to-severe asthma: pooled data from two randomised placebo-controlled studies. Thorax. 2015;70(8):748–56. https://doi.org/10.1136/thoraxjnl-2014-206719.

    Article  PubMed  Google Scholar 

  39. Ortega HG, Liu MC, Pavord ID, et al. Mepolizumab treatment in patients with severe eosinophilic asthma. N Engl J Med. 2014;371(13):1198–207. https://doi.org/10.1056/NEJMoa1403290.

    Article  CAS  PubMed  Google Scholar 

  40. Guilleminault L, Conde E, Reber LL. Pharmacological approaches to target type 2 cytokines in asthma. Pharmacol Ther. 2022. https://doi.org/10.1016/j.pharmthera.2022.108167.

    Article  PubMed  Google Scholar 

  41. Nedeva D, Kowal K, Mihaicuta S, et al. Epithelial alarmins: a new target to treat chronic respiratory diseases. Expert Rev Respir Med. 2023;17(9):773–86. https://doi.org/10.1080/17476348.2023.2262920.

    Article  CAS  PubMed  Google Scholar 

  42. Gauvreau GM, Hohlfeld JM, FitzGerald JM, et al. Inhaled anti-TSLP antibody fragment, ecleralimab, blocks responses to allergen in mild asthma. Eur Respir J. 2023;61(3). https://doi.org/10.1183/13993003.01193-2022.

  43. Luo W, Hu J, Xu W, Dong J. Distinct spatial and temporal roles for Th1, Th2, and Th17 cells in asthma. Front Immunol. 2022;13:974066. https://doi.org/10.3389/fimmu.2022.974066.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Newcomb DC, Peebles RS Jr. Th17-mediated inflammation in asthma. Curr Opin Immunol. 2013;25(6):755–60. https://doi.org/10.1016/j.coi.2013.08.002.

    Article  CAS  PubMed  Google Scholar 

  45. Simpson JL, Scott R, Boyle MJ, Gibson PG. Inflammatory subtypes in asthma: assessment and identification using induced sputum. Respirology. 2006;11(1):54–61. https://doi.org/10.1111/j.1440-1843.2006.00784.x.

    Article  PubMed  Google Scholar 

  46. Mizutani N, Nabe T, Yoshino S. IL-17A promotes the exacerbation of IL-33-induced airway hyperresponsiveness by enhancing neutrophilic inflammation via CXCR2 signaling in mice. J Immunol. 2014;192(4):1372–84. https://doi.org/10.4049/jimmunol.1301538.

    Article  CAS  PubMed  Google Scholar 

  47. Fossiez F, Djossou O, Chomarat P, et al. T cell interleukin-17 induces stromal cells to produce proinflammatory and hematopoietic cytokines. J Exp Med. 1996;183(6):2593–603. https://doi.org/10.1084/jem.183.6.2593.

    Article  CAS  PubMed  Google Scholar 

  48. Kolls JK, Lindén A. Interleukin-17 family members and inflammation. Immunity. 2004;21(4):467–76. https://doi.org/10.1016/j.immuni.2004.08.018.

  49. Wright TK, Gibson PG, Simpson JL, McDonald VM, Wood LG, Baines KJ. Neutrophil extracellular traps are associated with inflammation in chronic airway disease. Respirology. 2016;21(3):467–75. https://doi.org/10.1111/resp.12730.

    Article  PubMed  Google Scholar 

  50. Kudo M, Melton AC, Chen C, et al. IL-17A produced by αβ T cells drives airway hyper-responsiveness in mice and enhances mouse and human airway smooth muscle contraction. Nat Med. 2012;18(4):547–54. https://doi.org/10.1038/nm.2684.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Schnyder B, Schnyder-Candrian S. Dual role of Th17 cytokines, IL-17A,F, and IL-22 in allergic asthma. In: Quesniaux V, Ryffel B, Padova F, editors. IL-17, IL-22 and their producing cells: role in inflammation and autoimmunity. Basel: Springer Basel; 2013. pp. 143–55.

    Google Scholar 

  52. Leyva-Castillo JM, Yoon J, Geha RS. IL-22 promotes allergic airway inflammation in epicutaneously sensitized mice. J Allergy Clin Immunol. 2019;143(2):619–30. https://doi.org/10.1016/j.jaci.2018.05.032.

    Article  CAS  PubMed  Google Scholar 

  53. Nakagome K, Imamura M, Kawahata K, et al. High expression of IL-22 suppresses antigen-induced immune responses and eosinophilic airway inflammation via an IL-10-associated mechanism. J Immunol. 2011;187(10):5077–89. https://doi.org/10.4049/jimmunol.1001560.

    Article  CAS  PubMed  Google Scholar 

  54. O’Byrne PM, Metev H, Puu M, et al. Efficacy and safety of a CXCR2 antagonist, AZD5069, in patients with uncontrolled persistent asthma: a randomised, double-blind, placebo-controlled trial. Lancet Respir Med. 2016;4(10):797–806. https://doi.org/10.1016/s2213-2600(16)30227-2.

    Article  PubMed  Google Scholar 

  55. Clarke D, Damera G, Sukkar MB, Tliba O. Transcriptional regulation of cytokine function in airway smooth muscle cells. Pulm Pharmacol Ther. 2009;22(5):436–45. https://doi.org/10.1016/j.pupt.2009.04.003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Wisniewski JA, Muehling LM, Eccles JD, et al. T(H)1 signatures are present in the lower airways of children with severe asthma, regardless of allergic status. J Allergy Clin Immunol. 2018;141(6):2048–60. https://doi.org/10.1016/j.jaci.2017.08.020.

    Article  CAS  PubMed  Google Scholar 

  57. Britt RD Jr., Thompson MA, Sasse S, Pabelick CM, Gerber AN, Prakash YS. Th1 cytokines TNF-α and IFN-γ promote corticosteroid resistance in developing human airway smooth muscle. Am J Physiol Lung Cell Mol Physiol. 2019;316(1):L71–81. https://doi.org/10.1152/ajplung.00547.2017.

    Article  CAS  PubMed  Google Scholar 

  58. Namakanova OA, Gorshkova EA, Zvartsev RV, Nedospasov SA, Drutskaya MS, Gubernatorova EO. Therapeutic potential of combining IL-6 and TNF blockade in a mouse model of allergic asthma. Int J Mol Sci. 2022;23. https://doi.org/10.3390/ijms23073521.

  59. Mease PJ, Genovese MC, Greenwald MW, et al. Brodalumab, an anti-IL17RA monoclonal antibody, in psoriatic arthritis. N Engl J Med. 2014;370(24):2295–306. https://doi.org/10.1056/NEJMoa1315231.

    Article  CAS  PubMed  Google Scholar 

  60. Lee JH, Yu HH, Wang LC, Yang YH, Lin YT, Chiang BL. The levels of CD4 + CD25 + regulatory T cells in paediatric patients with allergic rhinitis and bronchial asthma. Clin Exp Immunol. 2007;148(1):53–63. https://doi.org/10.1111/j.1365-2249.2007.03329.x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Mamessier E, Nieves A, Lorec AM, et al. T-cell activation during exacerbations: a longitudinal study in refractory asthma. Allergy. 2008;63(9):1202–10. https://doi.org/10.1111/j.1398-9995.2008.01687.x.

    Article  CAS  PubMed  Google Scholar 

  62. Ling EM, Smith T, Nguyen XD, et al. Relation of CD4 + CD25 + regulatory T-cell suppression of allergen-driven T-cell activation to atopic status and expression of allergic disease. Lancet. 2004;363(9409):608–15. https://doi.org/10.1016/s0140-6736(04)15592-x.

    Article  CAS  PubMed  Google Scholar 

  63. Grindebacke H, Wing K, Andersson AC, Suri-Payer E, Rak S, Rudin A. Defective suppression of Th2 cytokines by CD4CD25 regulatory T cells in birch allergics during birch pollen season. Clin Exp Allergy. 2004;34(9):1364–72. https://doi.org/10.1111/j.1365-2222.2004.02067.x.

    Article  CAS  PubMed  Google Scholar 

  64. Kearley J, Barker JE, Robinson DS, Lloyd CM. Resolution of airway inflammation and hyperreactivity after in vivo transfer of CD4 + CD25 + regulatory T cells is interleukin 10 dependent. J Exp Med. 2005;202(11):1539–47. https://doi.org/10.1084/jem.20051166.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Kearley J, Robinson DS, Lloyd CM. CD4 + CD25 + regulatory T cells reverse established allergic airway inflammation and prevent airway remodeling. J Allergy Clin Immunol. 2008;122(3):617–24. https://doi.org/10.1016/j.jaci.2008.05.048.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Lan F, Zhang N, Bachert C, Zhang L. Stability of regulatory T cells in T helper 2-biased allergic airway diseases. Allergy. 2020;75(8):1918–26. https://doi.org/10.1111/all.14257.

    Article  PubMed  Google Scholar 

  67. McGee HS, Agrawal DK. Naturally occurring and inducible T-regulatory cells modulating immune response in allergic asthma. Am J Respir Crit Care Med. 2009;180(3):211–25. https://doi.org/10.1164/rccm.200809-1505OC.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Martín-Orozco E, Norte-Muñoz M, Martínez-García J. Regulatory T cells in Allergy and Asthma. Front Pediatr. 2017;5:117. https://doi.org/10.3389/fped.2017.00117.

    Article  PubMed  PubMed Central  Google Scholar 

  69. Li MO, Flavell RA. Contextual regulation of inflammation: a duet by transforming growth factor-beta and interleukin-10. Immunity. 2008;28(4):468–76. https://doi.org/10.1016/j.immuni.2008.03.003.

    Article  CAS  PubMed  Google Scholar 

  70. Burton OT, Noval Rivas M, Zhou JS, et al. Immunoglobulin E signal inhibition during allergen ingestion leads to reversal of established food allergy and induction of regulatory T cells. Immunity. 2014;41(1):141–51. https://doi.org/10.1016/j.immuni.2014.05.017.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Akdis M, Verhagen J, Taylor A, et al. Immune responses in healthy and allergic individuals are characterized by a fine balance between allergen-specific T regulatory 1 and T helper 2 cells. J Exp Med. 2004;199(11):1567–75. https://doi.org/10.1084/jem.20032058.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Akdis M, Akdis CA. Mechanisms of allergen-specific immunotherapy: multiple suppressor factors at work in immune tolerance to allergens. J Allergy Clin Immunol. 2014;133(3):621–31. https://doi.org/10.1016/j.jaci.2013.12.1088.

    Article  CAS  PubMed  Google Scholar 

  73. Vignola AM, Chanez P, Chiappara G, et al. Transforming growth factor-beta expression in mucosal biopsies in asthma and chronic bronchitis. Am J Respir Crit Care Med. 1997;156(2 Pt 1):591–9. https://doi.org/10.1164/ajrccm.156.2.9609066.

    Article  CAS  PubMed  Google Scholar 

  74. Boxall C, Holgate ST, Davies DE. The contribution of transforming growth factor-beta and epidermal growth factor signalling to airway remodelling in chronic asthma. Eur Respir J. 2006;27(1):208–29. https://doi.org/10.1183/09031936.06.00130004.

    Article  CAS  PubMed  Google Scholar 

  75. Wang W, Li P, Chen YF, Yang J. A potential immunopathogenic role for reduced IL-35 expression in allergic asthma. J Asthma. 2015;52(8):763–71. https://doi.org/10.3109/02770903.2015.1038390.

    Article  CAS  PubMed  Google Scholar 

  76. Collison LW, Workman CJ, Kuo TT, et al. The inhibitory cytokine IL-35 contributes to regulatory T-cell function. Nature. 2007;450(7169):566–9. https://doi.org/10.1038/nature06306.

    Article  CAS  PubMed  Google Scholar 

  77. Whitehead GS, Wilson RH, Nakano K, Burch LH, Nakano H, Cook DN. IL-35 production by inducible costimulator (ICOS)-positive regulatory T cells reverses established IL-17-dependent allergic airways disease. J Allergy Clin Immunol. 2012;129(1):207–e215. https://doi.org/10.1016/j.jaci.2011.08.009.

    Article  CAS  PubMed  Google Scholar 

  78. Zhou L, Lopes JE, Chong MM, et al. TGF-beta-induced Foxp3 inhibits T(H)17 cell differentiation by antagonizing RORgammat function. Nature. 2008;453(7192):236–40. https://doi.org/10.1038/nature06878.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Mantel PY, Kuipers H, Boyman O, et al. GATA3-driven Th2 responses inhibit TGF-beta1-induced FOXP3 expression and the formation of regulatory T cells. PLoS Biol. 2007;5(12):e329. https://doi.org/10.1371/journal.pbio.0050329.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Chen W, Cao Y, Zhong Y, Sun J, Dong J. The mechanisms of Effector Th cell responses contribute to Treg cell function: New insights into Pathogenesis and Therapy of Asthma. Front Immunol. 2022;13:862866. https://doi.org/10.3389/fimmu.2022.862866.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Jin H, Zhou Y, Wang L. The mechanism of rapamycin in promoting asthmatic regulatory T cell differentiation and function. Zhejiang Da Xue Xue Bao Yi Xue Ban. 2021;50(5):621–6. https://doi.org/10.3724/zdxbyxb-2021-0173.

    Article  PubMed  PubMed Central  Google Scholar 

  82. Pritam P, Manna S, Sahu A, et al. Eosinophil: a central player in modulating pathological complexity in asthma. Allergol Immunopathol (Madr). 2021;49(2):191–207. https://doi.org/10.15586/aei.v49i2.50.

    Article  PubMed  Google Scholar 

  83. Mukherjee M, Lacy P, Ueki S. Eosinophil Extracellular traps and Inflammatory pathologies-untangling the web! Front Immunol. 2018;9:2763. https://doi.org/10.3389/fimmu.2018.02763.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Choi Y, Kim YM, Lee HR, et al. Eosinophil extracellular traps activate type 2 innate lymphoid cells through stimulating airway epithelium in severe asthma. Allergy. 2020;75(1):95–103. https://doi.org/10.1111/all.13997.

    Article  CAS  PubMed  Google Scholar 

  85. Lambrecht BN, Persson EK, Hammad H. Myeloid cells in Asthma. Microbiol Spectr. 2017;5. https://doi.org/10.1128/microbiolspec.MCHD-0053-2016.

  86. Zhou X, Wei T, Cox CW, Jiang Y, Roche WR, Walls AF. Mast cell chymase impairs bronchial epithelium integrity by degrading cell junction molecules of epithelial cells. Allergy. 2019;74(7):1266–76. https://doi.org/10.1111/all.13666.

    Article  CAS  PubMed  Google Scholar 

  87. Carter RJ, Bradding P. The role of mast cells in the structural alterations of the airways as a potential mechanism in the pathogenesis of severe asthma. Curr Pharm Des. 2011;17(7):685–98. https://doi.org/10.2174/138161211795428975.

    Article  CAS  PubMed  Google Scholar 

  88. Elieh Ali Komi D, Bjermer L. Mast cell-mediated Orchestration of the Immune responses in human allergic asthma: current insights. Clin Rev Allergy Immunol. 2019;56(2):234–47. https://doi.org/10.1007/s12016-018-8720-1.

    Article  CAS  PubMed  Google Scholar 

  89. Dijkstra D, Meyer-Bahlburg A. Human basophils modulate plasma cell differentiation and maturation. J Immunol. 2017;198(1):229–38. https://doi.org/10.4049/jimmunol.1601144.

    Article  CAS  PubMed  Google Scholar 

  90. Holgate ST. New strategies with anti-IgE in allergic diseases. World Allergy Organ J. 2014;7(1):17. https://doi.org/10.1186/1939-4551-7-17.

  91. Matsui EC, Sampson HA, Bahnson HT, et al. Allergen-specific IgE as a biomarker of exposure plus sensitization in inner-city adolescents with asthma. Allergy. 2010;65(11):1414–22. https://doi.org/10.1111/j.1398-9995.2010.02412.x.

    Article  CAS  PubMed  Google Scholar 

  92. Burrows B, Martinez FD, Halonen M, Barbee RA, Cline MG. Association of asthma with serum IgE levels and skin-test reactivity to allergens. N Engl J Med. 1989;320(5):271–7. https://doi.org/10.1056/nejm198902023200502.

    Article  CAS  PubMed  Google Scholar 

  93. Mizutani N, Nabe T, Yoshino S. IgE/antigen-mediated enhancement of IgE production is a mechanism underlying the exacerbation of airway inflammation and remodelling in mice. Immunology. 2015;144(1):107–15. https://doi.org/10.1111/imm.12355.

    Article  CAS  PubMed  Google Scholar 

  94. Licari A, Marseglia G, Castagnoli R, Marseglia A, Ciprandi G. The discovery and development of omalizumab for the treatment of asthma. Expert Opin Drug Discov. 2015;10(9):1033–42. https://doi.org/10.1517/17460441.2015.1048220.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This declaration is not applicable.

Funding

This project is funded by The Natural Science Foundation of Jilin Province, project number: 20210101322JC.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed to the study conception and design. Yang Wang had the idea for the article and performed the literature search, and Li Liu critically revised the work. The first draft of the manuscript was written by Yang Wang and all authors commented on previous versions of the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Li Liu.

Ethics declarations

Conflict of interest

The authors have indicated they have no potential conflicts of interest to disclose.

Ethical approval and consent to participate

This declaration is not applicable.

Consent for publication

This declaration is not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, Y., Liu, L. Immunological factors, important players in the development of asthma. BMC Immunol 25, 50 (2024). https://doi.org/10.1186/s12865-024-00644-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12865-024-00644-w

Keywords